Skip to main content

Full text of "Introduction To Statistical Mechanics"

See other formats


a 

OU 166603 >m 



OSMANIA UNIVERSITY LIBRARY 

CallNo. pf / * * /\*\ fJL.- ^^^sion No. /(f X? 
Author 



This book should be returned on or before the date last marked belo 




INTRODUCTION TO STATISTICAL MECHANICS 

FOR STUDENTS OF PHYSICS AND 

PHYSICAL CHEMISTRY 



PUBLISHED BY 

Constable & Company Limited 
London W.C.2 



BOMBAY 

CALCUTTA MADRAS 
LEIPZIG 

Oxford, University 
Press 

TORONTO 

The Macmillan Company 
of Canada, Limited 



INTRODUCTION TO STATISTICAL 

MECHANICS'^ STUDENTS 

OF PHYSICS AND PHYSICAL 

CHEMISTRY 

BY 

JAMES EJUJi, M.A., 

ASSOCIATE PROFESSOR IN THE DEPARTMENT OF 
PHYSICS, UNIVERSITY OF LIVERPOOL. 

WITH A FOREWORD BY 

F. 0. DONNAN, C.B.E., M.A., Ph.D., D.Sc., F.B.S. 

PROFESSOR OK CHEMISTRY, UNIVERSITY COLLEGE, LONDON 



LONDON 

CONSTABLE & COMPANY LTD 
1930 



First published 1910 



PRINTED IN GREAT BRITAIN BY THE WH1TEFRIARS PRESS, LTD. 
LONDON AND TOJNBRIDGK. 



FOREWORD 

PHYSICO-CHEMICAL science regards the spatial universe as 
filled with a vast multitude of moving " elements " which 
possess both parfciculate and wave -like characters the 
" wavicles " of Eddington. The only way to render this 
elusive and protean microcosmos amenable to mathematical 
calculation, and to interpret physical measurement, is to 
deal in terms of probabilities, to employ the statistical 
method. The macroscopic world of sense and the measure- 
ments based thereon possess the validity characteristic of 
the averages of an actuarial estimate. The statistical method 
is therefore of profound importance. It dominates the 
whole of modern science. Combined with the generalised 
principles of dynamics and the quantum theory it has 
produced statistical mechanics and " quantum statistics." 
These are not special branches of science peculiar to the needs 
of a few lonely mathematicians. The simple truth is that 
they constitute the fundamental basis of modern physical 
science. But another simple truth, though a painful one, is 
that they involve a severe discipline for the untutored human 
mind. 

Fortunately, Professor Rice has now come to tutor our 
minds and bring consolation to our hearts. This book of 
his is a first-rate one, for which all serious students of 
chemistry and physics will owe him a deep and lasting debt of 
gratitude. He has explained and expounded the principles 
of statistical mechanics and quantum statistics with extreme 
lucidity. In the earlier portions of the book the general 
principles of probability and statistics are developed and 
applied to the solution of many important problems. Then 
the concepts of the quantum theory are introduced, and 
finally the generalised principles of dynamics. The most 
recent advances associated with the names of Bose, Einstein, 
Fermi and Dirac are dealt with in an important appendix. 
In excellent appendices to many of the chapters the author 



vi FOREWORD 

succeeds admirably in removing the mathematical difficul- 
ties inherent in parts of the reasoning. I would particularly 
and very warmly commend this book to the attention of 
students of physical chemistry it needs no recommendation 
to students of physics. Subjects such as chemical equilibria 
in gas reactions, the specific heats of gases and solids, the 
entropy of a perfect gas, the Nernst heat theorem, the 
chemical constants, the theory of the atom, the Einstciri- 
Smoluchowsky theory of density fluctuations and collision- 
form ula? arid chemical kinetics are all fully and clearly 
explained. Indeed, I will without hesitation make the two 
following assertions : (1) every student of physical chemistry 
must read the book ; (2) no student of physical chemistry 
familiar with the calculus will experience any serious diffi- 
culty in mastering its contents. 

I make these assertions because I believe this book is 
destined to exert a great influence on the training of the 
present generation of chemists and physicists. As the title 
indicates, Professor Rice has had in view and for that we 
cannot be too profoundly thankful the needs of chemists 
as well as physicists. He has certainly succeeded in his 
object. But he has done so without any slurring over or 
evasion of difficulties. Throughout the work the treatment 
is thorough and complete. I would specially commend to 
the attention of chemists (and physicists) the two chapters 
which deal in a most clear and original manner with the 
second law of thermodynamics. The light which this statis- 
tical analysis of the microcosmos throws on one of the greatest 
if not the greatest of the experiential generalisations 
of the macrooosmos constitutes, to my mind, one of the great 
triumphs of the human mind. Seldom, if ever, has it been set 
forth with such masterly lucidity and logic. 

This book is clearly the fruit of many years of study and 
thought. I wish it many years of prosperous and beneficent 
influence. 

F. G. DONNAN. 

THE SIR WILLIAM RAMSAY LABORATORIES OF 
PHYSICAL AND INORGANIC CHEMISTRY, 
UNIVERSITY COLLEGE, LONDON. 



PREFACE 

THE aim and scope of this book is indicated in the intro- 
ductory chapter. The author is not blind to the fact that 
the student he had in mind when he wrote it is not going to 
read some parts of it without a serious mental effort ; the 
necessarily mathematical form of the arguments entails that 
result. But he feels certain that any student with the 
average mathematical equipment acquired in the first two 
years of a University science course will not find it impossible 
to follow the details of the treatment, and as much assistance 
as possible is provided in explanatory appendices, in verbal 
interpretation and illustration. In the proofs no steps are- 
omitted. On that account the content of subjects dealt 
with has had to be restricted ; otherwise the book would 
have grown to a length unsuitable for the type of reader 
whom it is intended to assist. The problems treated are con- 
cerned with systems in statistical equilibrium, although a 
short appendix refers to the subject of collision-frequency in 
gases and its bearing on chemical reactions. 

The author's best thanks are due to Dr. A. McKeown 
and Mr. E. A. Stewardson, both of Liverpool University, for 
helpful advice and criticism, and to the former for reading 
the proofs. 

J. RICE. 

UNIVERSITY OF LIVERPOOL, 
October, 1929. 



Tii 



TABLE OF CONTENTS 

HAPTER PAGE 

FOREWORD ....... v 

PREFACE ....... vii 

INTRODUCTION ...... 1 

I. STATISTICAL METHOD ..... 6 

APPENDIX. THE NORMAL LAW OF ERRORS . 14 

II. THE STATISTICS OF A SIMPLE MOLECULAR SYSTEM 21 

APPENDIX A. STIRLING'S THEOREM . . 28 

APPENDIX B. PREPONDERANCE OF CERTAIN 

STATES ....... 29 

III. THE PROBABILITIES OF THE DIFFERENT STATES 

OF A SIMPLE MOLECULAR SYSTEM ... 34 

IV. TEMPERATURE AND THE DISTRIBUTION CONSTANT 45 
V. EXTENSION TO MORE COMPLEX MOLECULES . 57 

VI. THE SECOND LAW OF THERMODYNAMICS . . 68 
VII. THE ENTROPY OF A PERFECT GAS ... 77 
VIII. THE STATISTICAL THEORY OF CHEMICAL EQUI- 
LIBRIUM IN A GAS REACTION ... 82 
IX. INTERMOLECULAR FORCES .... 92 
X. FLUCTUATIONS OF DENSITY IN A MOLECULAR 

SYSTEM 102 

XI. THE SECOND LAW OF THERMODYNAMICS II . 114 
XII. THE STATISTICAL-MECHANICAL THEORY OF A 

LIQUID AND A VAPOUR PHASE IN CONTACT . 118 

XIII. THE SOLID STATE CONSIDERED AS A SIMPLE 

LATTICE OF MASSIVE PARTICLES . . .123 

XIV. THE QUANTUM HYPOTHESIS . . . .126 
XV. THE THEORY OF THE STATIONARY STATES OF AN 

ATOM 143 

XVI. DISTRIBUTION OF A SYSTEM IN ENERGY . . 151 
APPENDIX. DEGENERACY . . . .161 

S.M. ix b 



x TABLE OF CONTENTS 

CHAPTER PAGE 

XVII. QUANTUM THEORY OF THE SPECIFIC HEATS OF 

GASES 164 

XVIII. THE ELASTIC SPECTRUM OF A LINEAR LATTICE 

OF COHERING PARTICLES . . . .174 

XIX. THE ELASTIC SPECTRUM OF A CUBICAL LATTICE . 189 

XX. THE SPECIFIC HEATS OF SOLID BODIES . . 207 

XXI. THE ENTROPY CONSTANT OF A GAS . . 213 

XXII. THE ENTROPY CONSTANT OF A MON ATOMIC GAS 

AND STATISTICAL MECHANICS . . . 221 

XXIII. ENSEMBLES OF SYSTEMS I .... 234 

XXIV. ENSEMBLES OF SYSTEMS II .... 246 

APPENDIX ON RECENT DEVELOPMENTS 

I. BOSE'S STATISTICS OF LIGHT QUANTA IN A TEM- 
PERATURE-ENCLOSURE 270 

II. EINSTEIN'S THEORY OF AN IDEAL GAS . . 275 

III. THE FERMI-DIRAC STATISTICS .... 277 

IV. THE STATISTICAL METHOD OF DARWIN AND 

FOWLER 282 

APPENDIX ON COLLISION FORMULAE AND 
CHEMICAL KINETICS 

I. COLLISIONS BETWEEN MOLECULES IN A GAS . 296 
II. COLLISION-FREQUENCY AND EQUILIBRIUM. THE 

H THEOREM 306 

III. THE KINETICS OF GAS REACTIONS IN A HOMO- 
GENEOUS SYSTEM 311 

Note on Chapter X. THE EINSTEIN FLUCTUATION 

FORMULA ....... 327 

SUGGESTIONS FOR FURTHER READING . . 321) 

INDEX 331 



INTRODUCTION TO 

STATISTICAL MECHANICS FOR 

STUDENTS OF PHYSICS AND 

PHYSICAL CHEMISTRY 

INTRODUCTION 

THE experimental data which have acted as a guide to the 
discovery of the laws of physics and chemistry have been in 
the main derived from careful observation of the behaviour 
of very limited portions of matter deliberately and skilfully 
placed by the experimenter in artificial surroundings. The 
purpose of such an environment is the elimination as far as 
may be possible of all those external influences which under 
more normal circumstances would affect the behaviour of 
the portion of matter considered, but which are of no imme- 
diate interest to the observer, and only interfere with his 
search for the effect of some special influence with which he 
is at the moment directly concerned, and which is allowed by 
the special circumstances of the experiment to have full 
play. 

Such a portion of matter so situated may be termed a 
" system," and although its material constitution may not 
be simple, yet, by a strict limitation of the number of external 
influences operating on it and by a restriction of the number 
of properties observed, it may be possible to regard it as a 
system of a simple nature. By the removal of artificial 
limitations and the direction of the attention to a wider circle 
of properties the system becomes more and more complex. 
Thus a small quantity of dry air, confined in a glass tube 
above mercury, whose temperature is maintained constant, 
is an example of an extremely simple system whose change 
of volume under change of external pressure is the sole point 



2 STATISTICAL MECHANICS FOR STUDENTS 

of interest for the observer. The system would be equally 
simple if the pressure were maintained constant and observa- 
tion directed towards change of volume caused by change of 
temperature. The system becomes more complex if both 
pressure and temperature are allowed to vary, but the 
information already obtained from the simple cases enables 
us to predict (with success as it so happens in this case) what 
happens under the wider operation of external influences. 
If we now consider a portion of matter consisting of air and 
water, the system, now having two distinct phases (an air- 
vapour phase and a liquid phase), is more complex, inasmuch 
as we naturally observe two quantities, viz., the volume of 
each phase, when conditions of pressure and temperature 
are varied. Information derivable from the study of such 
a system and of others still more complex, is applied by the 
meteorologist to large tracts of our atmosphere or may even 
be of service in general considerations concerning the 
atmosphere as a whole regarded as a single but very complex 
system. 

Systems are sometimes defined as "physical " or " chemi- 
cal " according as the observations made are related to 
change in physical properties or in chemical constitution, 
but the terms are not always definite and no clear separa- 
tion into two classes is in general possible. We shall use 
the term " physical system " to embrace all systems in which 
chemical as well as physical properties are observed and to 
which the epithet " physico-chemical " might be attached 
save for its clumsiness as a word. 

Historically, the systems^hich were the first to be treated 
successfully by the methods of exact observation and 
mathematical analysis initiated in the sixteenth and seven- 
teenth centuries, are the mechanical devices, bodies moving 
on or near the earth's surface and the system of planets and 
satellites attendant on the sun. Here the observed proper- 
ties are the relative positions of the various bodies or parts 
of the system and the changes produced by the forces acting 
mutually between the various parts or exerted on them by 
bodies external to the system. The famous laws of motion 
propounded by Newton as an adequate summary of the 



INTRODUCTION 3 

experimental facts, and applied by him with wonderful skill 
and success to a wider range of such phenomena, were 
extended and provided with a very complete mathematical 
formulation by D'Alembert, Lagrange, Laplace, Hamilton 
and others. Such formulations became known as various 
analytical ways of stating the " Principles of Dynamics/' and 
a system of bodies in which the movement of its parts 
(assumed to conform exactly with these principles) is the 
prime object of observation is termed a " dynamical system." 
However, motion is only one of the observable features of 
any collection of bodies. Properties such as temperature, 
pressure, quantity of heat, luminosity, colour, refractivity, 
magnetic induction, electric charge and potential, chemical 
constitution, reactivity, etc., claim our attention. For some 
time several of these properties were explained by postulating 
the existence of subtle and weightless forms of matter, not 
accessible to direct observation, such as " caloric," " mag- 
netic and electric fluids." But the influence of the mathe- 
matical physics of the eighteenth century, with its treatment 
of the movement of a finite body as arising from the 
interaction of the discrete " particles " constituting the body 
and from their response to external forces, gave an irre- 
sistible impulse towards the explanation of all physical and 
chemical properties of matter as manifestations of the 
configuration and motion of the ultimate particles of the 
matter. It was not an accidental circumstance, but one 
quite natural in the mental environment of the time, that 
Dalton should have been led to the formulation of his atomic 
hypothesis by considerations of a mathematical-physical 
rather than of a purely chemical nature. There followed in 
quick succession the dynamical theory of heat, the kinetic 
theory of gases, the molecular theory of magnetism, all 
meeting with stubborn resistance but all winning recognition 
by their power in summarising experience and by the 
ultimate identity of the underlying ideas in each case. 
Finally, as the pinnacle of this edifice, built on a dynamical 
view of all properties of matter, there was constructed the 
theory that radiation is ultimately a propagation of an 
actual wave motion in a medium possessing elastic and 

B 2 



4 STATISTICAL MECHANICS FOR STUDENTS 

inertial properties like solid matter, those properties them- 
selves being thrown back in the writings of Cauchy and 
others on the interaction of ultimate ether particles far sur- 
passing material atoms in minuteness and fineness of struc- 
ture. Even biological science could not evade the influence 
of these ideas, and in the nineteenth century there was for a 
time a great vogue in the idea that in some way life and 
consciousness are but the by-products of the mechanical 
reactions between the ultimate atoms of living tissue, whose 
movements are just as much determined by the laws of 
dynamics as are those of the planets in our solar system. 
To be sure such crude notions have had their day in biology, 
and even in physical science, " atoms interacting across 
empty space " and " waves pulsating through the lumini- 
ferous ether " are strongly suspect. But although the 
conceptual entities which we employ in order to give our 
minds an orderly picture of the apparently inextricable 
complexity of natural phenomena are being replaced at the 
moment by new and as yet unfamiliar concepts, there still 
remains as powerful as ever and absolutely indispensable 
the great body of mathematical analysis which has grown 
up with the physical science of the past two or three cen- 
turies ; and in that body analytical dynamics holds a 
fundamental place for the reasons already stated. 

This mechanical conception of the underlying nature of 
physical phenomena is familiar enough to anyone conversant 
with the usual texts of Physics and Chemistry. The notions 
are entirely plausible in a qualitative or roughly quantitative 
sense. It is when one goes into the matter in some detail 
and attempts to apply mathematical methods in order to 
produce quantitatively precise or nearly precise results that 
trouble begins. Molecular systems are much too complex 
to follow in detail with the aid of dynamical laws and hypo- 
theses as to the nature of intermolecular forces or intra- 
molecular and intermolecular electromagnetic fields. To 
make any headway at all the laws of probability have to be 
impressed into service and made to co-operate with dynami- 
cal principles. The worker is involved at once in statistical 
considerations, and this combination of statistical calcula- 



INTRODUCTION 5 

tions with dynamical reasoning is called " Statistical 
Mechanics." Fortunately, for those not too conversant 
with dynamical methods, considerable progress in this 
subject can be made towards tangible results without any 
greater knowledge of mechanics than that possessed by the 
average student at the end of a University second year in 
the Applied Mathematics class room. At a pinch one can 
manage along for a time on even less. In this book every 
effort is made to keep at first to illustrations of such a nature 
that a detailed knowledge of dynamical methods, such as 
those employing the Lagrange and Hamilton equations of 
motion, is not required. For a satisfactory foundation, 
however, of the postulates upon which we base the statistical 
calculations, a knowledge of Hamilton's equations is required. 
Still, we shall assume that the postulates, explained at the 
outset, are all right, and make use of them at once, deferring 
their justification to the last chapters of the book. This 
appeals to the author as being probably the manner of 
laying out the work, which will evoke the interest of the 
reader at once. As regards pure statistics, little more is 
needed than an elementary knowledge of permutations, 
such as is available in any algebra text, and of the binomial 
and multinomial theorems. It is assumed, of course, that 
the reader has some knowledge of the symbolism and methods 
of the calculus. Any special mathematical information 
beyond this is supplied in appendices. 



CHAPTER I 

STATISTICAL METHOD 

1 . 1 The Spin of a Coin. It is a commonplace statement 
that on spinning a penny the chances are equal that it will 
present a head or tail. The d priori probability of either is 
0-5. The use of the epithet " a priori " might lead us to 
infer that this is a statement deduced from our " inner 
consciousness " or some equally mysterious source. Not 
so ; it is a bald statement of the experimental fact that if 
anyone chooses to amuse himself for some time by tossing 
a coin repeatedly and at random, he will find the ratio of 
heads to tails always close to unity, and the more so the 
longer he proceeds with the entertainment. There are two 
aspects presented by the fallen coin, and one has as good a 
chance of showing itself as the other. 

Suppose we spin two coins. How many " complexions " 
are possible ? There are four, since the two pennies may 
both present heads or both tails, or the first penny may 
present a head and the second a tail, or vice versd. Any one 
of these is as probable as any other, since the events are 
independent ; for the fall of one coin (say) head up, does not 
bias the fate of the other. Thus the a priori probability of 
each of the four complexions is 0-25, and any " doubting 
Thomas " can overcome his scepticism by trying it. He 
will find that on making a large number of throws 
practically one quarter of them will yield any given com- 
plexion. 

There are four complexions, but three " statistical 
states " : (1) both pennies showing heads, (2) both show- 
ing tails, (3) one showing a head and one a tail. Two 
complexions fall within state (3), and so the probability of 
that state is 0-5, while that of the states (1) and (2) are 0-25 
each. State (3) is twice as probable as either (1) or (2). 



STATISTICAL METHOD 7 

Increase the number of coins thrown to three. There are 
eight complexions. Here they are : 

First Penny. Second Penny. Third Penny. 

(1) ... Head ... Head ... Head 

(2) ... Head ... Head ... Tail 

(3) ... Head ... Tail ... Head 

(4) ... Head ... Tail ... Tail 

(5) ... Tail ... Head ... Head 

(6) ... Tail ... Head ... Tail 

(7) ... Tail ... Tail ... Head 

(8) ... Tail ... Tail ... Tail 

All are equally probable, having ^ as their a priori proba- 
bility. There are four statistical states, viz. : 

(1) All heads. 

(2) All tails. 

(3) Two heads and one tail. 

(4) One head and two tails. 

Only one complexion falls within either state (1) or (2), 
but three within either of the third and fourth states. 
Hence the states (3) and (4) are each thrice as probable as 
(1) or (2). The probabilities are J, , f , f . 

It should require little thought now to extend the reason- 
ing to any number of coins. Let there be n coins, and 
suppose we indicate the fact that the r th coin presents a head 
by the symbol a r , and that it presents a tail by the symbol b r . 
Then any complexion presented by a fall of the n coins will 
be represented by some such expression as 

aj a 2 6 3 a 4 6 5 a n . 

Each term will consist of n symbols. In every term, the 
suffixes will proceed regularly from 1 to n, but the arrange- 
ment of the a and b symbols will be fortuitous. Thus the 
term written represents a complexion in which the first 
penny falls head up, the second head up, the third tail up, 
the fourth head up, the fifth tail up, etc., the last head up. 
We can find all the possible complexions by working out the 
product of the n factors 

(l + &l) (2+^ 2 ) (8+*s) K + *n) (1.1-1) 

There are 2 n terms ; in each of them any suffix can only 



8 STATISTICAL MECHANICS FOR STUDENTS 

occur once as no penny can show both a head and a t^il in 
one fall. The b priori probability of each complexion is 
therefore 2~ n . How many statistical states are there and 
what is the probability of each ? A state, e.g., in which r 
pennies show heads and s show tails (r + s = n) will include 
all those complexions whose symbolic terms contain r of the 
a symbols and s of the 6 symbols. If we wish to find the 
number of these complexions we obliterate the suffixes in 
(1.1.1) and consider the coefficient of a r b* in the product 
(a + b) n . By the binomial theorem this is 

n ! 
TIT! 

Each of the complexions has an a priori probability 2~ w . 
Hence the probability of the statistical state mentioned is 



r \s\ \2J 

There are, of course, as many statistical states as there 
are terms in the expansion of (a -f b) n , i.e., n + 1. 

The reader should bear in mind the experimental basis 
of these calculations. The results might have been other- 
wise. If by some edict of the Master of the Mint pennies 
were so loaded that each one fell twice as often head up as 
tail up, the d priori probability of a head would be 2/3, of a 
tail, 1/3. To take account of the increased chance of a head 
being shown by any coin we must in the reckoning of com- 
plexions consider the product 

(2a x + 6 X ) (2a 2 + 6 2 ) (2a n + b n ). 

The coefficient of any term will give the number of times 
which the corresponding complexion will turn up on the 
average out of 3 n tosses. Thus a complexion in which r 
particular pennies jburn up heads and the remaining s tails 
will have an b priori probability of (2/3) r (l/3)< or 2 r /3 n . So 
the probability of the statistical state, r heads and s tails 
would no longer be the coefficient of a r b 9 in 



STATISTICAL METHOD 
but in 



In general, if the a priori probabilities of a head and a tail 
were respectively # and q (p + q = 1), the probability of the 
statistical state mentioned would be the coefficient of 
a r b' in 

(pa + q b)*, 
i.e., 

FTTi^ * ' ' ' (1 ' 1 ' 3) 

1 . 2 Throwing o! Dice. If we were to use dice instead of 
coins, we should have six possibilities with each die, not 
merely two as in the case of coins. Each aspect of a die 
has an a priori probability 1/6, if the dice are not loaded. 
Let a throw of one by the r ih die be symbolised by a lf , of 
two by a 2r , etc., of six by a 6f . We can symbolise a given 
complexion in which, say, the first die throws a three, the 
second die a five, the third a five, the fourth a one, the fifth 
a two, the sixth a four, the seventh a three, etc. the n ih a 
two by the expression 



In any such symbolic term the second suffix advances regu- 
larly from 1 to n, but the first is fortuitously chosen from 1 
to 6. Any possible complexion is represented by some one 
of the 6 n terms obtained by expanding the product of the 
n factors. 

fall + 021 + + a 6l) (012 + <*22 + + #62) 

...... (*+* + . + 0eJ (1-2.1) 

Each complexion has an a priori probability of 6~ w . 

We can find the number of different complexions within 
the statistical state, in which n l dice throw a one, n 2 dice 
throw a two, etc., 7^ 6 dice throw a six, by eradicating the 
second suffix in (1 . 2 . 1) and calculating the coefficient of the 
term. 



a s n 



10 STATISTICAL MECHANICS FOR STUDENTS 
in the expansion of 

(i + 2 + a 3 + a * + a s + a e) n 

In any text-book of algebra the reader will find in the 
chapter on the multinomial theorem that this is 

n\ 



Thus the probability of this statistical state is 

n\ 



\ n \ n I n \ n \ n ! V6 



(1.2.2) 



If the dice had been loaded alike, so that the a priori prob- 
ability of a throw of a one by any die were p l9 of a two, p 2 , 
etc., (#! + p% + Pz + Pt + Pz + P Q ~ I)> the probability 
of the state mentioned would be the coefficient of a/ 1 a 2 w * 
a 3 n a 4 w * a 6 w ' a 6 n in 

i.e. 9 

V f 

IV 9 p *i p *t p * p n* p * p * f (1.2.3) 

Wj! n 2 \ %! n 4 ! n 5 \ n^l ' l 2 3 4 5 6 

The extension of these results is now an easy matter. 
Let there be n similar articles each one of which can present 
at one time one of c different aspects, the a priori probability 

of each aspect being p l9 p 29 p& , p c respectively, then 

the probability of the statistical state in which n articles 
present the first aspect, n 2 the second aspect, etc., n c the 
c th aspect is 



P.*' 



//it/i/it / i * * * * 

n^n^. n c \ 

the first factor being the number of complexions consistent 
with these aspects. 

The total number of possible complexions, being the 
number of separate terms in a product of n factors each 
containing c terms, viz., 

(a n + a 21 + . . . + a cl ) (a 12 + a 22 + . . . + a c2 ) . . . 



STATISTICAL METHOD 11 

is, of course, c n . On the other hand, the number of statistical 
states is the number of terms in the expansion 

K + a 2 + ....+ a.)* 

which is, therefore, the number of " homogeneous products " 
of the c quantities a l9 a 2 , . . . , a c , each product being of the 
7i th degree. Reference to a text-book of Algebra will show 
that this is 

(' + *- 1 )' .... (1.2.6) 

V 



1 . 3 The Normal Law of Errors. We shall now consider a 
modification of this random throwing of a number of articles 
which will lead to the introduction of a mathematical 
function which at a later stage plays an important part in 
statistical-mechanical reasoning. 

Instead of coins or dice, let us have in our possession n 
counters, each one being labelled + e on one side and 
on the other. On throwing these, any complexion will, if 
we add the numbers showing, yield a sum me where m is a 
positive or negative integer lying between n and + n. 
In fact m is r s where r counters show positive faces and 
s counters show negative ; thus the probability of the sum 
me is given by the number of complexions corresponding to 
(r, s). It is 

" .... (1.3.1) 

* ; 



rls 
where 

r -j- s n 
r s = m 

By a well known result, the expression ( 1 . 3 . 1 ) is maximum 
when r = s, and the value decreases progressively to the 
amount (\) n as the difference r s or s r increases in 
numerical value from zero to n. Thus the state in which the 
sum is zero is the most probable, and if n is a very large 
number, the probability of those states in which the sum is 
zero or only a small multiple of e, far outweighs the prob- 
ability of those in which the sum takes a relatively large 
numerical value. It will be both interesting and serviceable 



12 STATISTICAL MECHANICS FOR STUDENTS 

to investigate the limiting form for this expression (1.3.1) 
when is made very small in value and the number n grows 
without limit. This condition is, however, rather vague, 
for it makes no provision for the maximum value of the 
sum, viz., nc. We can make the condition sufficiently 
precise by postulating that ne may also increase without 
limit as e decreases and n increases, but in such a manner 
that the product of the maximum value of the sum, nc, 
and the common difference, 2e, between consecutive values 
of the sum remains finite and constant. Thus write 



where k is a finite constant. So as not to interrupt the 
general course of the reasoning we relegate some rather 
tedious mathematical steps to an appendix where it is shown 
that in the limit the expression (1 . 3 . 1) is equal to 



where z is written for (r - s) e and 8z = 2e. 

Thus it appears that the probability that the sum lies 
between the values i\ and <T 2 , is given by the expression 



A exp(-=-,)dz . . . (1.3.3) 




It is a well known result that the value of the integral in 
(1.3.3) between the limits oo and + oo is kn*, and so the 
expression (1.3.3) when Cj is oo and C a is + is unity, 
as, of course, it must be since the sum of the probabilities of 
all possible states must be unity. 

The integral 



plays an important part in statistical-mechanical analysis, 
as we shall see later, and a number of its most useful pro- 
perties are summarised in the appendix. But before passing 

* The series 

1 + L + 2! + j -f . . . . + -f . . . . ad. inf. v 
11 2! 31 n! Jy 

is written exp (y). It is, of course, equal to ev, if y is a real quantity. 



STATISTICAL METHOD 13 

on to the application of the results of this chapter to mole- 
cular systems, it may be as well to point out that we have 
in this section been treating an important case in the Theory 
of Errors of Observations. 

We may assume that the accidental error in an obser- 
vation made with a definite instrument is the algebraic sum 
of a number of component errors of the instrument due to 
change in external conditions, uncorrected errors of the 
instrument and peculiarities of the observer. Each of these 
components may in its turn be regarded as due to a large 
number of elementary causes, so that we are not violating 
any obvious truth in assuming that an accidental error of 
observation is the algebraic sum of a very great number of 
very small errors. Suppose we make a further simplifying 
assumption (which can only be justified by subsequent 
comparison of the results of making it with the facts) that 
all these elementary errors have the same numerical magni- 
tude, but are as likely to be positive as negative. Thus if 
there are n elementary errors altogether, each of magnitude 
e, the actual error will in any observation have the magnitude 
(r s)e if r of the elementary errors are positive and s are 
negative in that observation. But such an outcome will 
have the same probability as in the case of the counters, 
where r came down positive and s negative. That is, the 
probability of the error (r s)e is given, by (1 . 3 . 1), and 
proceeding as before we find that if the true value of the 
quantity observed is a, the chance that an observation gives 
a value lying between a + C i and a + f 2 * s indicated by the 
expression (1.3.3). But those interested may refer to the 
appendix for further information. It is more to our purpose 
to proceed to the study of molecular systems, and the reader 
can glance at the appendix when at a later stage he is com- 
pelled to know something about the integral. 



14 STATISTICAL MECHANICS FOR STUDENTS 
APPENDIX TO CHAPTER I 

ON THE NORMAL LAWS OF ERRORS 

THE problem raised in section (1 . 3) is to obtain a function 
of a continuous variable z which shall replace 

n\ 



-(-V 

r! \2/ 



this being a function of a quantity (r s)e which varies by 
discrete amounts 2e. Implicitly we have somewhat changed 
our point of view. Previously an error which was not an 
integral multiple of c, was not supposed to occur at all ; its 
probability was zero ; probabilities were, so to speak, con- 
centrated on definite errors. Now all errors are possible 
from z = oo to z = + oo, and, what may appear para- 
doxical at first sight, the probability that the error may 
have any definite value is zero. But this is only natural as 
the number of possible errors now is unlimited, and the 
chance for any one of them is one in infinity, or nothing at 
all. The form of the statement must now be that there 
exists a function of z,/(z), such that the probability that the 
error shall lie between the values z = d and z = C 2 will be 
given by the integral 



rf. 

/(*) d*, 
J & 



and to find /(z) we must proceed on the assumption that 
there is an approximate equality between 

/(z)> 
and 

n\ /l\ n 
H7T\2J 

where z == (r s)e and 8z = 2e, the approximation being 
closer and closer the smaller e, and therefore the larger r and" $ 
(and, of course, n) for a given value of z. The reader is 
warned against a too common misconception that /(z) is 
the chance of an error z. He is asked to bear in mind that 



STATISTICAL METHOD 15 

the function which replaces (n \ / r ! 8 !) . 2~ tt is not/(z), but 
f(z) Sz. The infinitesimal is as important a factor of the 
function as/(z) itself. The probability which was previously 
concentrated on a distinct value of z is now as it were spread 
over a small neighbouring range of values. 
Writing m for r 8, we have 



"*">*,-& o 

2e/(me + 2e) ~ 

^ ' " 



and also 



r + Us - 1! 
Hence 

f(m e + 2 6) ^ a 

/(m e ) " r + 1 
and so 

/(me + 2 e) -/(me) _._ r - s + 1 ^_ __ m + 1 



/(me + 2c 
or 
/(m e + 2 e) -/(m c) _.__ m + 1 



26 2(71 + l) 

On writing z for me, 2 for 2e and k 2 for 2ne 2 , and pro- 
ceeding to the limit, we obtain 



This is equivalent to 

d log /(z) _ 



Hence 



= + constant, 

iC 



or 



The integration constant A is to be determined by the fact 
that 



16 STATISTICAL MECHANICS FOR STUDENTS 

since the integral represents the sum of all the probabilities 
and must be unity. 

Before proceeding it may be as well to state briefly a 
number of results concerning the integral 



and kindred integrals. These will prove serviceable in later 
chapters. By a simple change of variables, x = za*, we can 
write for this integral 

a -j| &-** fa. 

First of all it can be shown that if the integral is taken 
between the limits x = oo and x = + oo, we obtain 

(^ * 

1 e x dx =TT 

J ~ co 

Since e~ x * is an even function of x t it is also true 

e~ x * dx = -77*. 

^ o 

The value of 

~[ e~ x *dx 



is a function of 77, which increases from zero to unity as 77 
increases from zero to infinity. Tables of this function for 
definite values of 77 are printed in text-books on the Theory 
of Errors.* If 77 is equal to 0-1 the expression is about 
0-1125 ; if 77 is equal to 0-5, it is 0-5205 ; for 77 equal to 1, 
the value is 0-8427 ; and by the time 77 has reached 3, 
the expression has attained the value 0-99998 ; so that the 
integration from 3 to oo contributes only -00002 of the value 
to the function of 77. One can realise the truth of this 
in a general way by noting that if x = 0-1, e~ x * = 

* See for example, Combinations of Observations, Brunt, A table is 
also printed in Jeans' Dynamical Theory of Gases. 



STATISTICAL METHOD 17 

0-99905, while if x = 3, er* = -00012, and beyond this 
e~ x * decreases to very minute values with great rapidity. 
It can also be shown that 



if the index n is an odd integer, and 

/oo 

x n e~ x ' dx = *, 



if ft is an even integer. 

The cases of this which we shall require now and at a later 
stage are 

co 

x e-*' dx = ~ 



[ \ 

x* e~ x * dx = ~ 

I 4 

/ 



x 2 e~ x * dx = - 



We can now determine the constant A in the equality 
above 



A p-02* 

( A e , 



f(z)=Aexp (_. 

\ /C 

A p 02* 

A 6 , 

where we write a for I/A 2 . For if 



18 STATISTICAL MECHANICS FOR STUDENTS 

then 



since 



Thus the chance that an error may fall between the values 

z = x and 2 = C 2 is 





This law has been deduced with the aid of rather restrictive 
assumptions concerning the nature of the causes giving rise 
to errors. Actually, rather broader assumptions can form 
a starting point for its deductions, and it is found to be 
closely followed in many cases. Of course in some circum- 
stances other error laws hold ; e.g., if the restriction that 
positive elementary errors are as likely as negative be 
removed, we cannot arrive at a function symmetrical in 
value with respect to the origin. 

The reader will naturally inquire as to the part played by 
the constant k (or a) in the law. The way in which it was 
introduced (as equal to e(2?i)*), gives no clear indication of 
this ; but we can easily arrive at a conclusion concerning it 
by inquiring into the average error made in a large series of 
observations which conform to the normal law. The equal 
preponderance of positive and negative elementary errors 
shows that the average of the observations is the true value 
of the quantity observed, denoted by a. The chance that 
an observation lies between a + z and a -\- z + 8z is 
(a/7r)* e~ a * f Sz. Now as negative and positive values of the 
error z are equally likely, the average error in the strict 
algebraic sense is zero ; but this is of no help to us. However, 
we can find the average numerical value of the errors if we 
multiply the numerical value of z by its probability (a/7r)* 
er** 82, and sum over the whole range, i.e., integrate from 
Oto oo. 



STATISTICAL METHOD 19 



7T/ a 
1 



Another important manner of estimating an average 
value for the error is to square each error, multiply this by 
the probability in each case, and sum over all the errors. 
This result gives us the mean square of the errors, and the 
square root of it is called the " mean square error " (M.S.E.). 
Thus we find 



-*' dz 




= *! 

2 

and so the M.S.E. is &/2*, and the ratio of the average 
numerical error to the M.S.E. is (2/7r)* or 0-798. 

Still another result concerns what is called the " median 
error," which has such a value that if we denote it by r, 
then half the errors lie between r and + r. This quantity 
is determined by the equality 



02 




20 STATISTICAL MECHANICS FOR STUDENTS 



or - =0-5. 




From the tables of 1 e""*" dx, it can be found that m* has 

JG 

the value 0-4770, or r = -4770 A;. 

It is clear then that the value of the constant k indicates 
the standard of precision of the measurements made. The 
smaller k is the more accurate has been the series of 
observations. Of course all this implies that care and skill 
have already been expended on the actual observational 
work. No amount of tinkering with the theory of errors is 
going to draw reliable conclusions from careless measure- 
ments. 



CHAPTER II 

THE STATISTICS OF A SIMPLE MOLECULAR SYSTEM 

2 . 1 The Complexions of a Molecular System. Configura- 
tions. We begin our statistical-mechanical investigations 
by applying the methods developed in the previous chapter 
to a simple body which is conceived to be an aggregate of 
molecules, each molecule being regarded as having no 
structure, but merely possessing minute mass and extension 
in short the ' ( particle " of dynamical theory. Any obvious 
changes in the body must, of course, be associated with 
unobservable but yet actual changes in the relative con- 
figuration and motions of the particles ; but, to be sure, a 
great deal of change might be going on in the relative situa- 
tions and movements of the molecules without any observable 
difference manifesting itself in the general appearance and 
behaviour of the body. 

The configuration of the molecular system can only be 
precisely defined by assigning definite co-ordinates to each 
particle (the particles for the moment being regarded as 
equivalent to points), and as the configurations possible for 
a given volume of the body are therefore unlimited in number 
it is impossible to deal with statistical states by counting 
the number of configurations consistent with each state, 
since these are uncountable. In the Theory of Errors we 
regard an error as having a calculable chance if it lies in an 
assigned range of errors ; so in the example before us we 
regard a complexion of the molecules as defined by assigning 
a range of configuration. In short, we do not say that a 
particular molecule is at a definite point (x, y, 2), but that 
its co-ordinates lie between x and x + 8x, y and y + 8y, z 
and z + Sz where Sx, Sy, 8z are assigned small quantities. 
In other words, we divide the volume occupied by the body 
into a, finite number of " cells," and state that such and such 

21 



22 STATISTICAL MECHANICS FOR STUDENTS 

molecules are at the moment situated in such and such cells. 
As far as the counting of complexions is concerned, we 
regard the molecules as analogous to the coins or dice of the 
previous chapter. The existence of a molecule in an assigned 
cell is analogous to the exhibition of a certain aspect by a coin 
or die. So long as a molecule remains in the same cell, its 
" aspect " is unchanged. A " complexion " of the system 
is defined by the particular aspects shown by the individual 
molecules, i.e., by the way the individual molecules are 
distributed among the cells. Merely to move the molecules 
about within the cells does not alter the complexion ; but 
the transfer of molecules from cell to cell will alter the com- 
plexion even although it is only an interchange which leaves 
the number (but not the individuality) of the molecules in 
the cells unaltered. The smaller the cells and the greater 
their number, the finer is the detail, so to speak, which 
distinguishes one complexion from another, but in order to 
render counting conceivably possible, the size of each cell 
must remain finite and the number of them also finite, 
though possibly very large. We shall see later that in 
carrying on the mathematical analysis after the counting 
has been effected, we may have to resort to integrations 
which imply infinitesimal elements of volume unlimited in 
number in much the same way as in Chapter I. we passed 
from summation over a range of discrete numbers to integra- 
tion throughout a range of a continuous variable ; but in 
the initial stages of the analysis the assumption of finite 
cells is necessary for beginning the calculation at all. 

If, therefore, there are n molecules and c cells, we find just 
as in section (1.2) that the number of complexions in which 
n v molecules are in the first cell, n 2 in the second cell, etc., 
...... , n c in the c th cell is 

- _J^ --- . . . (2.1.1) 



We shall in future denote this number by the functional 
symbol 

W (n lt 2> ....... a ). 

All this is, of course, a mere matter of counting. Many of 



STATISTICS OF A SIMPLE MOLECULAR SYSTEM 23 

these complexions would entail the crowding of the body 
into one small portion of the volume within its external 
surface. Indeed there are c complexions in which all the 
molecules would be in one cell. But this consideration need 
not deter us from proceeding ; for if we make the proviso 
that the cells though small enough to make c a large number, 
are yet large enough to contain a large number of molecules 
at the average density, so that in fact the ratio of n to c is 
also large, then it is possible to show that in all but a 
relatively small number of -complexions the molecular 
density is uniform or sufficiently near uniformity for the 
discrepancy to be undetectable by experimental means. It 
is, of course, easy to see that W (n ly n 2 , ...... , n c ) is in- 

creased in value if in the denominator two numbers, n r and 
n 8 , in any pair of factorials are replaced by two other 
integers which have the same sum, but are more nearly 
equal to one another. 
For 

(k + l)\(k l)l 
= (k + l) (4+Z- 1) ...... (k + l).k\x 

k\ 



>k\k\ 
and therefore 



_ 

(k + l)\(k-l)\ k\k\ 

Thus the state in which the density is uniform embraces a 
number of complexions which is greater than for any other 
state in which the numbers n l9 n 2 , ...... ,n e are given, but 

are not all equal to n/c.* But to prove the statement con- 
cerning the preponderance in number of the complexions 
corresponding to uniform density or states near it over all 
other complexions requires closer analysis than this ; and 
to proceed with the proof we require the assistance of an 
approximation to W (n lt n 2 , ...... , n c ), which, however, 

* We suppose the number of molecules and number of cells to be so 
chosen that n/c is an integer. 



24 STATISTICAL MECHANICS FOR STUDENTS 

has to be used with due consideration for the conditions 
under which it is true. 

There is a famous theorem published early in the eighteenth 
century by the Scottish mathematician James Stirling, 
which states that there is an approximate equality between 
the logarithm (to the Napierian base) of the factorial of n 
and the expression ' 

(n + 2 j log, n - n + ~ log, 2 IT. 

Provided n is as large as 10, for example, a four-figure 
logarithm table will not reveal the discrepancy, and for 
sufficiently large numbers, it is possible to write 

log n \ = n log n - n. 

(See the appendix for some remarks on Stirling's theorem.) 
From this it follows that, provided none of the numbers 
n r are too small, 

c 

log W (n l9 n 2> , n c ) = n log n 2 n r log n r (2.1.2) 

r=i 

With the aid of this approximation one can prove the 
assertion made above. The details of the proof will be 
found in the appendix to this chapter, and the outcome is to 
demonstrate as we have already said, that the statistical 
states in which the molecules are uniformly distributed, or 
distributed in such a manner that the number in each cell 
differs but little from the average, embrace all but a negli- 
gible fraction of the total complexions. The point of this 
important result will appear presently. 

2 . 2 The Complexions of a Molecular System. Phases. 
In the previous section we confined our attention to the 
arrangements of the molecules in space ; but, of course, in 
any attempt to explain the general properties of matter by 
dynamical theory, we must also take account of the motions 
of the molecules relative to the frame of reference in which 
the body as a whole is regarded as fixed. Since the molecules 
are regarded as particles, it is easy to conceive a graphical 
representation of a precise velocity-condition. Choosing an 
origin, the velocity of a particle can be represented by the 



STATISTICS OF A SIMPLE MOLECULAR SYSTEM 25 

position of a point in a " velocity-diagram, " the vector from 
the origin to the point representing the magnitude and 
direction of the velocity. Thus the actual velocities of the 
n particles at any time are represented by a configuration in 
the velocity -diagram of the n representative points. This 
diagram can, of course, be divided into cells as in the case 
of the volume occupied by the body ; velocity-complexions 
can be defined just as before and the number of complexions 
embraced in any velocity-state (where velocities within 
narrow limits are assigned without reference to the indi- 
viduality of the molecules) can be counted and comparisons 
of relative numerical strength be made. 

It is customary to link up the two methods of partitioning 
into one " picture." The configuration and velocity of each 
molecule is' regarded as an entity with six components, three 
position-components and three velocity-components, and is 
called a " phase " of the molecule. When precise values of 
velocity and position are assigned to each molecule in the 
system, we are said to have prescribed a " phase of the 
system/' As before, the phases possible to the system are 
unlimited in number, and no progress can be made in 
counting complexions unless a complexion is regarded not 
as a phase, but an arrangement of the molecules in a small 
but finite " extension-in-phase." That is, we do not desig- 
nate an " aspect " of a molecule by saying that its position 
and velocity are given by components x, y, z, u, v, w, but by 
saying that the first co-ordinate lies between x and x + Sx, 
etc. ; the first velocity-component between u and u + 8u 

etc., where Sx, , Sw are finite but small increments. 

We can visualise as small rectangular figures the forms to 
which we attach the symbols, and regard 8x By Sz and 
Su 8v Sw as their volumes. With no possibility of visualising, 
we nevertheless refer to the magnitude Sx Sy Sz Su Sv Sw as 
an element of " extension-in-phase," and for convenience 
and brevity call it a " phase-cell," borrowing the geometric 
term from our earlier considerations. Indeed, in a great 
deal of the literature of the subject, geometrical language is 
used in a manner which at the outset may dismay the 
beginner, who imagines he is called upon to perform the 



26 STATISTICAL MECHANICS FOR STUDENTS 

feat of " seeing " a space of six or more dimensions. He 
may reassure himself ; no such impossible task is expected 
from him. The geometric terms are merely borrowed and 
attached to analogous ideas ; indeed their use can be avoided 
altogether (e.g., that is the practice of Gibbs), but for the 
sake of reading the general literature later, the beginner 
should familiarise himself with this use of geometric terms. 
For instance, a phase of a molecule which is indicated by 
assigning definite values to the x, y, z, u, v, w, of the molecule, 
is referred to as a " point in the six-dimensional phase- 
diagram/' An extension in phase, which is a range of values 
of position and velocity co-ordinates such that no phase in 
the range has co-ordinate values outside the six ranges x to 
x -\- 8x, w to w + Sw, is called a cell of the phase- 
diagram. We can conceivably visualise the path of a mole- 
cule in physical space. It is the geometric counterpart of a 
continuous series of values of the position co-ordinates. We 
can also visualise a curve in a velocity-diagram which would 
represent geometrically the changing values of u, v, w, for a 
molecule. Indeed the reader has perhaps met it in his 
academic text-books of mechanics under the name " hodo- 
graph." When we place the two pictures together, we 
cannot get a " picture," but we still use the geometric 
language ; the series of continuous phases through which 
the position and motion of a molecule pass with lapse of 
time, is called the " phase-path " or " trajectory " of the 
molecule. 

If the reader still experiences any difficulty in this matter 
of geometric terminology, he may find the following device 
of some assistance, until he becomes sufficiently familiar to 
dispense with it. Let him think of three plane diagrams, 
each one provided with the usual pair of rectangular axes. 
In one let x and u be represented by a point ; in the second 
let y and v be so represented ; and in the third, z and w. A 
phase of a molecule can then be visualised by thinking of 
three points one in each diagram ; an extension in phase 
by thinking of three small rectangles ; the changing phases 
of a molecule by thinking of three curves. A " point-group " 
represents a phase ; a " curve-group " represents the history 



STATISTICS OF A SIMPLE MOLECULAR SYSTEM 27 

of the molecule and a " rectangle-group " represents an 
extension-in-phase. The division of the phase-diagram into 
cells may be associated in the mind's eye with the cross- 
meshing of the plane diagrams by lines parallel to the axes. 
One has to be careful, however. If even one member of a 
point-group is changed, the phase is altered ; we have 
changed to a different point in the phase-diagram. Keeping 
two rectangles in the group alike, but changing the third, 
produces a different phase-cell. Suppose, for example, we 
have divided a limited portion of the (x, u) diagram into I 
rectangles, another limited portion of the (y, v) diagram 
into m rectangles, and of the (z, w) into n rectangles, we are 
dealing with a limited extension in the phase-diagram which 
has been divided into I m n cells ; for any rectangle in the 
(#, u) diagram can be combined with any in the (y, v), and 
this again with any in the (2, w), each combination pro- 
ducing a distinct phase-cell. 

The algebraic work from this point is just as before. If 
there are c phase-cells and n molecules, the number of com- 
plexions embraced in that state in which the phases of n l 
molecules are within the first cell, etc., of n c molecules 
within the c th cell is W (n l ,n 2 , n e ) 

where 

W (i' > ' ra c) = ,_, rr 

'H- n 2' U C- 

and approximately 

c 

log W (n l9 rc 2 , n c ) = n log n 2 ri r log n r . 

r=l 

So far there has been no reference to dynamical con- 
siderations. We have been concentrating our attention on 
the concepts necessary for the statistical side of the work. 
We must now turn our attention to the second member of 
the double barrelled epithet, " statistical-mechanical/' 



28 STATISTICAL MECHANICS FOR STUDENTS 



APPENDIX TO CHAPTER II 

A. Stirling's Theorem. A rigorous proof of this theorem, 
which states that 



can only be appreciated by those who have given some pains 
to the study of series and their convergency properties. 
Such a proof will be found by those interested in ChrystaFs 
Algebra, Vol. II. Chap. XXX. For a more general theorem 
which takes the place of Stirling's when n is not an integer, 
a reference can be made to Whittaker and Watson's Modern 
Analysis, Chap. XII. 

For the majority of his readers, the author suspects that 
a simple empirical test will be quite satisfactory. If any 
one cares to take the trouble to look up Napierian logarithms 
in a book of tables (the cheap little book prepared by C. G. 
Knott will suffice), he will find that for instance 

log 5 - ^g 1 + log 2 + ...... + log 5 

* 5 

has the value 0*6519. He will also find that 

log 10 + log 77 



1 - 



10 



has the value 0*6553. If he chooses a larger number 10, he 
will find that 

l og 10 - lQ g * + lQ g 2 + + lQ g 10 

^ 10 

has the value 0-7922 ; and that 

- 1 __ log 20 + log 77 
20 

is also 0*7922, so that a four-figure table cannot distinguish 
between the values of the two expressions. This empirical 
procedure gives considerable support to the general result 
that 



STATISTICS OF A SIMPLE MOLECULAR SYSTEM 29 
* + lo 2 + ...... + l n 



log n 



ft 
_., log 2 ft 4- log 



2ft 
or 

log ft ! == ft log ft ft 4~ - log (2 77 ft). 

J 

It follows that 



.e 

A rather more general (but still far from rigorous) pro- 
cedure is to begin from the well-known integration theorem 

I log x dx = x log x x -\- constant. 
Taking this between the limits 1 to ft, we have 

fn 

I log x dx = ft log ft ft 4~ ! 

We can replace the integral by the approximate expression 
(x 2 x ) log x + fc 3 # 2 ) log# 2 + . . . 4- (x f x f __i) log x f _ l9 

where x v x 2 , x s , , x f are a series of equally spaced 

values of x ranging from 1 to ft , provided any of the differences 
x r x r _ l are small compared to the x r . If ft is very large, 
we can makes these differences unity, and thus replace the 
integral by 

log 1 4- log 2 4- + log ft. 

Thus approximately 



n 



E log r === ft log ft ft -f 1 
i 

= ft log ft ft. 

B. The Preponderance of Certain States as Regards the 
Numbers of Complexions Embraced in Them. We have seen 
that in a state in which the distribution of the molecules 
among the cells is n v n 2 , ...... , n c , the " complexion- 

number " W (ft 1? ft 2 , ...... , ft c ) is given by the approxi- 

mation 

e 

log W (n l9 ft a , ...... n c ) = ft log ft 2 n r log ft, . 

r-l 



30 STATISTICAL MECHANICS FOR STUDENTS 

We shall denote the expression on the right-hand side by 
k (n v n 2 , ...... , n c ). The function k (n lt n 2 , ...... , n c ) 

has its maximum value when all the n r are each equal to 
n/c \ this value we shall denote by k m . So the maximum 
value of W (n v n 2 , ...... n c ) is W m where 

lo g w = k m = n (log n log a), 
a being written for n/c. 

To investigate the relative numerical strength of states 
in the neighbourhood of this state, let us write 

n 1 =a+p l ,n 2 =a+ P 2 , ...... , n c = a + &, 

where p l9 /? 2 , ...... , p c are a set of positive or negative 

integers which must satisfy the relation 



r-1 

It follows that k (n v n 2 , ...... n c ) for this state is con- 

nected with k m by the equation 

k m -k=Z\(a+p r )log(a+p r )\ -nloga 



using the well-known expansion of log (1 + x). 

If we work out this expression we obtain a series whose 
terms are multiples of the expressions 2 /? f , 27 $, 2 , S r 3 , 
etc. The first term is, of course, zero, and we find after a 
little rearrangement that 



-Tn^l) +etc ' 
10r=i\ay 

For small values of the /? f , the series reduces to its first term, 
which, being a sum of squares, is essentially positive, as we 



STATISTICS OF A SIMPLE MOLECULAR SYSTEM 31 

should naturally expect since k m > k for any values of the 

& 

In the text of the chapter, the statement was made that 

the number of complexions embraced in these states for 
which the jS r are zero or equal to small fractions of a, are 
sufficiently great to " swamp " the complexions embraced 
in states diverging in a perceptible degree from uniformity 
of distribution. The justification of this depends on equation 
(I.). On its right-hand side the ratio of the second term to 
the first is 

1 

3 a 

and if the ratios of the various fi r to a are sufficiently small, 
this is also small, so that the second and also the subsequent 
terms of the series can be neglected, and we can write 



___ ^ __ y r 2 

K -^ ^ x r * 



where 



We thus arrive at the result that if W is the complexion 
number for the state defined by the integral values f3 l9 j8 2 , 

j8 c for deviation from uniformity, then W satisfies 

approximately the equality. 

W = e* 



aR*\ 
J 



where 



W, of course, only attaining the maximum value W w where 
every j3 r (or x r ) is zero. 

Now let us for the moment turn our attention to the follow- 
ing little problem. Despite its apparent irrelevance, it will 
soon be apparent why we do so. 



32 STATISTICAL MECHANICS FOR STUDENTS 

Imagine that we have a series of small particles dotted 
about in space at the points of a regular cubical lattice, and 
that the mass of any particle is equal to m e~ ar * where r 
is the distance of the particle from an origin which is itself 
a point of the lattice, m being the mass of the particle at the 
origin. If we were required to find the total mass within a 
sphere of given radius Z, and the elementary cubes of the 
lattice were small enough, a sufficiently accurate answer to 
the question would be found by supposing the matter to be 
distributed continuously and not at discrete points, so that 
the density at a point is given by p e~ ar \ p being the 
density at the origin. The mass within a sphere I would be 



r 2 e~ ar dr, 

J 

which is equal to 



f 1 
r 

Jo 



where 

= m* 



Now we saw in the appendix to the last chapter that the 
definite integral 



is equal to ?r*/4 ; it is also true that the whole of this value 
is practically contributed between the limits to 3 ; which 
shows that in our problem almost the entire mass of the 
lattice of particles is within a sphere whose radius is three 
times a""*. 

To return, after this digression, to the main argument, 
the solution of the problem concerning the preponderance 
of complexions is just a multidimensional analogue of the 
lattice problem with an analogous answer. A state can be 

supposed to be represented by a " point " x v x%, , x c 

in a c-dimensional space. At the " point " we suppose 
some entity of magnitude W located. We can then, in order 
to reduce the mathematical process to an integration, 
suppose the entity distributed uniformly with a density 



STATISTICS OF A SIMPLE MOLECULAR SYSTEM 33 



A C 

A exp [ - 

P \ 2 

and the amount of the entity within a certain extension of 
the c-dimensional space can be determined by the integral 

A I ...... \ e~ ~*~ dx l dx 2 ...... dx c 

throughout the extension. 

The knowledge of integration required is rather outside 
the usual academic courses delivered to students for whom 
this book is intended. Those interested will find the neces- 
sary material under the heading " Dirichlet's Integrals " in 
the more advanced texts on mathematical analysis, and the 
full working out of this problem in Jeans' Dynamical Theory 
of Gases. The upshot is similar to that of our lattice problem. 
Practically the whole W-magnitude of the entity is within 
an extension close round the origin, that is where none of 
the x r exceed rather small integral multiples of (2/a)*. 
Since a is assumed to be a large number, this implies that 
practically all the complexions possible to the molecular 
system are concentrated in states for which the j8 r are small 
compared with a, i.e., states in which deviation from 
uniformity is not marked. 

It is true that in the above the approximations which we 
have made exclude implicitly from the analysis states where 
there are only a few molecules in some cells, since the par- 
ticular use of Stirling's theorem is rather wide of the mark 
for small values of n r . However, a simple consideration of 
the original factorial expressions will show how relatively 
unimportant such states are ; and in any case one can use 
the still closer approximation for log n\, viz., (n + ^) 
log n n + ^ log 2 TT, and find the conclusion still justified, 
although the expressions are a little more complicated. 



CHAPTER III 

THE PROBABILITIES OF THE DIFFERENT STATES OF A 
SIMPLE MOLECULAR SYSTEM 

3 . 1 The a priori Probability of a Complexion. So far we 
have been confining our attention to the counting of com- 
plexions. In order to discover the probability of a state, 
we clearly must know or postulate something about the 
probabilities of the complexions included in it. But, it must 
be admitted that here we are faced with one of the most 
difficult questions in the whole subject. There is nothing 
novel, however, in such a situation. A satisfactory settle- 
ment of the postulates of any of the mathematical sciences 
fih^i proved to be one of the most difficult tasks, calling for 
intellectual endowment of no mean order on the part of 
those who have been the pioneers in this elusive branch of 
knowledge. Many a young student acquires a really skilful 
mathematical technique during his school and university 
years, without being aware of the doubtful nature of some 
of the processes employed unless hedged round with an 
array of carefully-phrased conditions. Not that there is in 
this state of affairs anything calling for severe censure on 
the part of the educational reformer. It is doubtful whether 
much good can come from too much immersion in " founda- 
tions " on the part of a mind as yet immature. No sane 
teacher ever dreams of troubling boys and girls with closely- 
reasoned disquisitions on the postulates of geometry. 
Certain statements very plausible to the young concerning 
congruence, parallelism and the like, are accepted at the 
outset, and a start is made on the deduction of theorems, 
arousing immediate and practical interest in the minds of 
the pupils. 

The author feels that the situation at this stage of our 
subject is essentially similar. He surmises that his readers 

34 



PROBABILITIES OF STATES 35 

are more anxious to " see results " as soon as possible, 
without being worried at the moment with a logical dis- 
cussion of postulates, which, if they have any element of 
plausibility about them, will be accepted without demur, 
especially if they lead to conclusions in agreement with 
experimental fact. Such a posteriori justification may 
indeed be entirely satisfactory for many ; but of course it 
would not be sound to leave matters in such a condition, 
and some attempt will be made at a later stage to discuss 
the basis of the postulates used. This, however, cannot be 
done until the mathematical expression of dynamical laws 
has been carried a stage further than is usual in the ele- 
mentary texts of dynamics. 

The fundamental postulate introduced at this stage is 
that if the various phase-cells have the same magnitude, 
one aspect of a given molecule is as probable as any other ; 
i.e., it is as likely for its " representative point " in the 
" phase-diagram " to be in any phase-cell as in any other. 
There is a certain plausibility about the assumption, which, 
however, must not blind us to the fact that some time or 
other it should be subjected to careful scrutiny. Clearly the 
history of the system is determined by dynamical law, and 
if our postulate be true, it must at least be a possibility that 
any molecule can in a sufficiently long lapse of time have 
been in every cell. In the first chapter the postulate of 
equal a priori probability of each aspect for an article after 
a throw was based on the fact that repeated throwing does 
show every aspect for an article in approximately equal 
numbers during a long time. No such direct experimental 
evidence comes to our assistance here ; we are not on such 
familiar terms with individual molecules as with coins and 
dice ; nor have we the services of " Maxwell demons " to 
call upon ! However, let us make the assumption for the 
sake of practical progress now and worry about it later. 

Certain obvious reservations must be made in the appli- 
cation of this postulate, however, which call for immediate 
notice. If the body of which our molecular system is regarded 
as an analogue, be solid, we find it hard to admit that any 
molecule can be in any " configuration-cell " into which we 

D 2 



36 STATISTICAL MECHANICS FOR STUDENTS 

divide the volume, even in long lapses of time. No doubt 
\ve can appeal to the experiments of Roberts -Austen and 
others on diffusion in the solid state ; but still the essence 
of the molecular picture of a solid is that the molecules, 
even if they are not fixed in unchanging neighbouring 
positions relative to one another, are vibrating about the 
points of some lattice as origins. And if each molecule 
cannot wander from configuration-cell to configuration-cell 
at random, neither can its representative point in the phase- 
diagram roam fortuitously about the phase -cells. This 
difficulty, however, is not serious. It serves to show that 
we shall have to treat the solid state on a somewhat different 
plan to the fluid states where our postulate has a greater air 
of plausibility. We shall find that in the solid state the 
interest centres round the vibrations about the set of fixed 
points and the co-ordinates and velocities of each molecule 
with reference to its own individual " origin " are the 
quantities which determine the phase at any definite instant, 
and it is with regard to partitions of a phase-diagram con- 
structed on such an understanding that our postulate will 
be introduced. 

For the present, therefore, we confine ourselves to the 
fluid states, and in order to simplify our preliminary con- 
siderations we shall begin with the gaseous state where no 
difficulties will be raised by intermolecular forces. In short, 
we shall at the outset deal with a monatomic gas. 

Another condition which restricts our postulate concerns 
the question of energy. In partitioning into configuration- 
cells, a natural limit is placed on the cells of a configuration- 
diagram by the external surface of the body. In a velocity- 
diagram no such boundary is obvious, yet dynamical con- 
siderations yield one very readily. If a velocity-cell were 
sufficiently far from the origin, the velocity represented by 
its central point" might be so great that if even a single 
molecule were possessed of this velocity, its kinetic energy 
would be greater than the energy-content of the body con- 
sidered. Thus a sphere of definite radius excludes velocity- 
cells which cannot come within our consideration. Indeed, 
instead of appealing to a hard, impenetrable boundary to 



PROBABILITIES OF STATES 37 

limit the configuration-cells, we might adopt a similar 
method of limitation as for the velocity -cells. That is, 
conceive that the region occupied by the gas is a region in 
which a field of force exists whose potential is vanishingly 
small unless one approaches the boundary, where it rises 
rapidly as we proceed outwards to values so great, that were 
even one molecule to be in this external region, its potential 
energy would exceed the energy-content of the body. Thus 
we place a natural limit on the phase-cells, to which the 
postulate of equal a priori probability can be applied, by 
means of one condition, viz., the definite energy-content of 
the body of which the molecular system is a model. 

One other point should be mentioned, before proceeding 
to definite probability calculations. For a reason which 
may not appear necessary at present, but which will be 
recognised as a matter of great convenience later, we will 
assume that the components of the phase of a molecule will 
include not the components of its velocity, but of its 
momentum. Of course, at present, this amounts to nothing 
more than a change of scale in the phase diagram ; but the 
change will justify itself very decidedly when dealing with 
more complex systems. 

If there are c cells of equal magnitude in the phase- 
diagram, the probability of a molecular representative point 
being in one of them, is by the postulate of equal probability 

equal to c" 1 ; i.e., in the formula (1 . 2 . 4)^9 1 ~p% 

= p c = c" 1 , and the probability of the state (n ly n 2 , 

T& C ), would appear to be 

W (n lt n 2 , ,n c ) c~ n . 

But this overlooks the limitation placed upon the total 
number of possible complexions by energy considerations. 
If the body of which the molecular system is a model is 
considered to be at constant temperature, the energy-content 
is given, and the total number of complexions is not c w , for 
many of these would be inconsistent with the equations 

n l + n 2 + + n c = n . . . (3.1.1) 

n l l + n 2 2 + +tt c c = E. . . (3.1.2) 



38 STATISTICAL MECHANICS FOR STUDENTS 

where 19 e 2 , , e c are the energies of a molecule when 

its representative point is in the first, second, , c th 

cell respectively and E is the energy-content of the body. 
We have to consider the sets of positive integers n l9 n 2 , 

,n c which satisfy (3. 1 . l)and(3. 1 .2), arid add together 

the W (n lt n 2 , n c ) functions for each set. If the sum 

is s, then s is the total number of possible complexions, and 
the probability of the state (n l9 n 2) , n c ) is 

W (n v ?2 2 , . . .,n c ) 

s ... (3.1.3) 

These considerations clearly take account of the condition 
mentioned above which determines the boundary of the 
possible phase cells ; for any cells for which e > E are 
excluded automatically by (3.1.2). The quantities e r are 
the energies corresponding to the centres of the cells. They 
will be given by some expression such as 

e -i- ^ + c 2 , , , , 

- 2 +^ (w) ' 

where , ??, Care the components of momentum for the phase, 
m the mass of the molecule, and $ (x, y, z) the potential 
energy of the molecule in the phase, due to any external 
field of force, such as gravity. For certain phase-cells on the 
boundary <f> (x, y, z) will also include the potential energy, 
mentioned earlier, which has an evanescent value in the 
main portion of the body's volume, but becomes rapidly 
significant as we approach the external surface, rising to a 
value greater than E. In the functional form of </) (x, y, z) 
will occur certain coefficients entering as factors of the 
powers and products of x, y, z in the various terms or in 
some other recognised manner. E.g., the quantity g will 
occur if gravity is regarded as acting on the molecules. 
These quantities are called " parameters/' They are 
regarded as constants in investigations on the probabilities 
of given states for a system with given energy, but, as will 
appear later, when drawing thermodynamic conclusions 
from statistical mechanics we have to consider changes in 
energy -content, and one way of producing such change is 
by an alteration of parameters, such as is caused by a change 



PROBABILITIES OF STATES 39 

in external bodies producing a field of force or in a move- 
ment of the external surface of the body ; this, in fact, being 
the analogue of " external work " done by or on the system. 
The reader is reminded that we are at the moment dealing 
with a gaseous system where intermolecular forces play no 
part. If it were not so, the energy of the system would 
include the mutual potential energy of the molecules which 
would not be proportional to the first power of the con- 
centration and in the expression on the left side of (3.1.2) 
the r could not be regarded as merely functions of the phases 
and parameters, but would also depend on the values of 

n l9 n 2 , , n c . In other words, E could not be put 

equal to a linear function of the n r . This consideration must 
seriously modify the treatment of liquids as compared with 
that of gases. 

The expression (3.1. 3) clearly implies an equal probability 
for each of the s possible complexions, and we refer once 
more to the necessity for a closer scrutiny at a later stage 
of the postulated foundations. This is no matter of repeated 
random " throwing " of molecules into cells. In casting 
dice or coins they are gathered up and thrown in a manner 
subject to no law other than the " law of chance." But 
even if we imagine that the molecular system is just now in 
any complexion we like, the subsequent complexions of its 
history do not follow " by chance," but are the results of 
movements and collisions subject to the law r s of dynamics. 
Waiving, however, closer investigation of this knotty point 
for the present, we proceed to discover the state with the 
maximum probability. 

3 . 2 The State of Maximum Probability. This will no 

longer be given by the equality of n v n 2 , , n c . The 

condition (3 . 1 .2) alters the whole character of the solution. 
As before, we have to find the values of the n r which will 

make W (n ly n 2 , , n c ) maximum (for given E, s is 

also given), but subject to the condition (3.1.2) as well as 
(3.1.1). Postponing for the moment the details of the 
solution, it appears that in the most probable state n l = v^ 

n% = v 2 , , etc., where 

v r =Ce~^ . . . . (3.2.1) 



40 STATISTICAL MECHANICS FOR STUDENTS 

C and JJL being two quantities which are functions of E, n 
and the parameters. These are determined by the two equa- 
tions (which are, of course, conditions (3.1.1) and (3.1.2) 
for this state), 



c 



2 e-^r = n . . . (3.2.2) 



c 



C S r e-^'r =E . . . (3.2.3) 

r 1 

Division of (3 . 2 . 3) by (3 . 2 . 2) gives 
Z e r er^r E 



2 e' 1 "' n 



(3.2. 4). 



Equation (3.2.4) determines /x as a function of the average 
molecular energy E//& and the parameters, and thereupon 
(3 . 2 . 2) or (3 . 2 . 3) will determine C. 

When we set out the steps of the solution presently we 
shall see that, since it follows the usual procedure of deter- 
mining maxima and minima in the calculus, we are treating 
the n r as continuous variables for the time being and, of 
course, there is no guarantee that any of the expressions 
C* e~"^r are integral. This is one of those minor troubles 
which beset us when we are engaged in discussing bodies 
with molecular structure by a mathematical method which 
practically implies that we are dealing with a continuous 
medium. The consequences, as we shall see, are in certain 
connections too serious to be overlooked and will compel us 
to adopt some form of " quantum hypothesis/' but in the 
present connection we are after all dealing in the main with 
such large values of n r that a change of unity while not 
exactly infinitesimal is relatively so small that the results 
of introducing the infinitesimal variations of the calculus is 
not going to lead to serious error. In (3 . 2 . 1) we can regard 
v r as such an integer that C e~* f r lies between and v r and 
v r 1, if not exactly equal to v r . The form of the solution 
once more excludes certain cells without any appeal to 
physical boundaries ; for if e r is large enough, e~^ e f is so 
small that C e~^ f f is a proper fraction. It follows that in 
(3.2.2) and (3.2.3) the limits of the summations may be 



PROBABILITIES OF STATES 41 

omitted and the series regarded as infinite since beyond a 
certain term there will be a negligible residue. 

A proof exactly on the lines of that in the appendix to 
Chapter II. can be constructed to show that if we consider 
the probabilities of states given by 

n r = r + & 

then the combined probabilities of the most probable state 
and those states for which the /3 r have relatively small 
values practically swamp the probabilities of all other states. 
Thus, subject to a satisfactory settlement of the doubtful postulate 
of equal probability of individual complexions in the prolonged 
history of the gas, we find that there is a " normal state " of 
the molecular model in or near which it will always be, 
except for brief and insignificant intervals of time. It is 
this state which we clearly must investigate if we are to 
derive by statistical-mechanical methods the well known 
thermodynamic properties of a system in thermodynamic 
equilibrium. We shall begin this task in the next chapter, 
and conclude this one by laying out the proof of the result 
(3.2.1). 

As before we seek the " max-min " condition for log W 
regarded as given by the approximation 

c 

n log n Z n r log n r . . . (3.2.5) 

r = l 

subject to the conditions (3.1.1) and (3.1. 2). As already 
stated we can without serious error regard the n r as con- 
tinuous variables. The method employed is known as the 
" Method of Undetermined Multipliers," and although not 
absolutely necessary, a perusal of an exposition of the 
method in a text of the calculus will prove helpful to any 
reader not familiar with it. 

Let us alter the n r to n r -f- Sn r , then the variation in 
log W is given by 

8 log W = - S (1 + log n r ) Sn r . . (3.2.6) 
and the variations, Sn r , must, on account of (3 . 1 . 1) and 
(3.1.2) satisfy the two equations 

ZSn r = . . . . (3.2.7) 
Z r 8n r = . . . . (3.2.8). 



42 STATISTICAL MECHANICS FOR STUDENTS 

But if the values of n r be such as to make log W maximum 
or minimum for any small variations from these values, then 
the equation 

27 (1 + log n r ) Sn, = 0, 

or, by reason of (3 . 2 . 7), 

Z log n r Sn r = . . . . (3.2.9) 

must be true as well as the equations (3.2.7) and (3.2.8). 
If A and /z are any quantities whatever, it follows as a matter 
of course that 

Z (log n r + A + /*,) 8n r = . . (3.2.10) 
This being so, let us choose X and /z so that 

log n + A + fi l = 
log n 2 + A + p, 2 = 

This is quite possible ; these are two simple simultaneous 
equations determining the two quantities A and /z uniquely 
as functions of l9 e 2 , n l and n 2 ; in fact 

_ lo^gj- log n^ 
^ ~ i *z 
A =-*i. Iog -^ 2 Jr_ .2 .l?gi 

1 2 

It follows that with these two values of A and \i 
(log n z + A + ^e 3 ) 8^3 + (log n + A 



etc. =-0 . (3.2.11) 

This result must be true for any values of the variations 
8n 3 , 8n 4 , ...... ; for having chosen any set of them, we can 

choose 8n l and 8n 2 to satisfy 



=0. 



But this cannot T3e so unless the individual multipliers of 
Sft 3 , 87i 4 , etc., in (3 . 2 . 11) are themselves zero, provided, of 
course, that A and //, have the values given above. Thus it 
transpires that in the condition when log W, and therefore 
W/#, has a maximum or minimum value, the following 
equations are all true 



PROBABILITIES OF STATES 43 

. (3.2.12) 



log n 


i+A+/x 


l =Q 


log?i 


2 + A + p 


e 2 = . 


log n 


3 + A + p 


r = 









log n r + A + /i r = 

These equations determine the n f as functions of the c r 
and A and /z, and if these values are inserted in (3 . 1 . 1) and 
(3 . 1 .2) they determine A and JJL as functions of the c r , E and n ; 
and so determine the n r as functions of the e r , E and n. In 
short, c equations, such as (3 . 2 . 12) and the equations (3.1.1) 
and (3.1. 2), constitute c + 2 simultaneous equations deter- 
mining uniquely the c + 2 quantities n l9 n 29 n c , A, p, 

in terms of the e r , E and w. The result is 

n r C e~^v, 

where C e~\ and as pointed out above, (3.2.4) will then 
determine p, and (3 . 2 . 2) or (3 . 2 . 3) will determine C or A. 
Since the equations determining C (or A) and p, involve 
summations over all the cells, which in practice will amount 
to integrations throughout the region of the phase -diagram 
bounded by the energy-condition, C and p, will not depend 
on individual e f , i.e., on individual phases, but will be 
functions of the parameters which, in conjunction with the 
phase, enter into the functional form giving any e f in terms 
of , 77,, f r , x r9 y r , z r . 

We still have to satisfy ourselves that the corresponding 
value of W is a maximum and not a minimum. This is 
almost obvious " on sight," but if any one requires a formal 
proof, he has only to make a small modification of the 
analysis in Appendix B to Chapter II. Writing v r + /?, for 
v r , we find 

log W m - log W - Z{(v r + j8 r ) log (v r + p r ) - v r log v r ] 

(and after a few steps on precisely similar lines to those in 
the appendix) 



2 



]} 



44 STATISTICAL MECHANICS FOR STUDENTS 

For small values of the /3 r the right-hand side is positive, 
no matter what the sign of the /3 r (for, of course, the v r are 
positive). 

Hence log W TO > log W, and W w is therefore a maximum 
and not a minimum. A continuation of the discussion leads 
as before to the enormous preponderance of the combined 
probabilities of the normal state and states near it over the 
combined probabilities of the remaining states. 



CHAPTER IV 

TEMPERATURE AND THE DISTRIBUTION CONSTANT 

4 . 1 The Average Energy of a Molecule. When we begin 
to consider some practicable method of calculating p and C 
from equations (3.2.2) and (3 . 2 . 3), it is clear that no pro- 
gress is possible by methods of series summation unless 
some definite information is available concerning the magni- 
tude of the phase-cell. There is nothing in the so-called 
" classical " dynamical methods to give us any help in this 
respect ; in fact it is one of the signal benefits which the 
quantum hypothesis has conferred on our mathematical 
methods that it has in conjunction with experimental work 
on spectroscopic phenomena suggested an answer to this 
difficulty. However, we shall have to defer this matter to 
a later stage. In the meantime we shall have to convert the 
series summations into integrations, thus leaving in our 
expressions an entirely undetermined quantity, which, for 
many purposes, is no drawback. Thus we shall have to 
replace a symbol such as n r by an integral 



J ..... J 



?> C,x,y,z)dd7]d(dxdydz y 



the integration being extended over the r th phase-cell, 
/(> ??> C ^) y> z) being some continuous function of the 
variables , 77, C, #, y, z. The solution for the state of 
maximum probability worked out in the last chapter shows 
that for v r we must write the above integral with the function 
/ (, ^, C x, y, z) given the form 

- (p + ^ + (*)-[*<(> (x 9 y, z) 
zm 

where D is a constant. 
Confining ourselves for the moment to the case in which 

45 



46 STATISTICAL MECHANICS FOR STUDENTS 
the gas is free from external force, it follows that v r is replaced 

by 

D[ ..... (exp { -a ( z + 7j 2 + P)}d{dT) dtdxdydz (4.1. 1)* 

extended over the r th phase-cell, where a = /*/2m. Further 
the quantity e r v r must be replaced by 

+ ^ + C 2 ) exp { - a\P + ^ + C 2 ) ! 

d ..... dz (4.1.2) 
also extended over the r th cell. 

The equations (3.2.2) and (3.2.3) are now replaced by 



~~ a + j = n 



" a 



(4.1.4) 

where v is the volume of the system. The summations 
were over all phase-cells, so that the integrations are now 
practically from oo to co for each variable. Using the 
results quoted in the appendix to Chapter I. we obtain 
from (4.1.3) 

. Dt; (~\ = n . . . . (4.1.5). 
To deal with (4,1.4) we first observe that 

dC 



-i (5)' ever 



* Of course t will involve x, y, 2, in the boundary cells where the 
limiting field of force acts ; but since is very large in these, no practical 
contribution to the integral is made by such cells. 



TEMPERATURE AND DISTRIBUTION CONSTANT 47 
and so (4.1.4) simplifies to 

?^(*Y = E .... (4.1.6) 
2m 2 \a 5 / V ' 

Hence, dividing (4 . 1 . 6) by (4 . 1 . 5), we obtain 

3 _ E 
4ma n 

3n . . . . (4.1.7) 

""as 

and, in consequence 

D = 

v 

n (4 1 8) 

' ' I*' 1 ' 5 ' 



The result (4.1.7) shows that the ''distribution-constant " 
fj, of the system is inversely proportional to the average 
energy of a molecule, so that if the temperature of the 
system rises, the distribution of the molecules among the 
phase-cells in the normal state is altered since //, decreases. 
Thus there is a direct relation between p and temperature, 
and it can be readily obtained by considering the pressure 
of the gas. 

4 . 2 The Equation of State of a Perfect Gas. The con- 
nection between the pressure of the gas and the velocities of 
its molecules is a simple one which requires for its derivation 
no other assumption than the postulate that the directions 
of the molecular velocities shall be distributed uniformly 
between all possible directions. A molecule which crosses 
an element of surface within the gas with a velocity u in a 
direction making an angle <f> with the surface, transfers 
across the surface an amount m u cos < of momentum normal 
to the surface. Let N (u, <f>) stand for the number of mole- 
cules per unit volume which have velocities within infinitesi- 
mal limits of u, and directions of motion within infinitesimal 
limits of the direction <f>. Of such molecules those that cross 
the element of surface in unit time lie in a volume 

A u cos <f> 



48 STATISTICAL MECHANICS FOR STUDENTS 

where A is the element of area. Such molecules therefore 
transfer normal momentum at the rate 

A N (w, 0) m u 2 cos 2 <f> 

Summing this for all values of (/>, and remembering that 
the average value of cos 2 ^ is 1/3,* we find that this rate of 
transfer of normal momentum across A due to molecules 
with a velocity u or infinitesimally near it, is 

- A N (u) m u 2 

where N (u) is the number of molecules per unit volume 
having velocities u or near it. 
This is 



where e(^) is the kinetic energy \ m u 2 of a molecule. Sum- 
ming for all molecules, we find that the rate of transference 
of normal momentum across the surface A is 

2 E 

\ . 

3 v 
Thus the pressure is 

2 E 

3 ~v 

or two-thirds of the energy-density.! Hence if p is the 

pressure 

2 

pv =_ E 
1 3 

by(4.1.7) =* ..... (4.2.1). 

P 

But the well-known equation of state is 

pv = R0 ..... (4.2.2) 

* If a, /?, 7 are the three direction-cosines of any direction with respect 
to three rectangular axes, 

cos 2 a 4- cos 2 j3 -f cos 2 7 = 1, 
and so _ . _ _ ^ 

cos 2 a = cos 2 =B cos 2 7 =-=. 
o 

t Further remarks on this method of calculating the pressure will be 
made in section (4.5). 



TEMPERATURE AND DISTRIBUTION CONSTANT 49 

where 6 is the absolute temperature and R the gas-constant 
for the given quantity of gas. Thus a comparison of (4 . 2 . 1) 
and (4.2.2) yields 



and since R is proportional to n (for at the same temperature 
and pressure the number of molecules is proportional to the 
volume), we have 

p." 1 =k9 ...... (4.2.3) 

where 

7 R 

k = 

n 

and is called the " gas-constant per molecule." 

The simple connection between the distribution-constant 
and the temperature of the gas is now apparent. Incidentally 
by (4 . 1 . 7) 

-=-&0 ..... (4.2.4) 
n 2 v ' 

or the average energy of a molecule is 1-5 times kd. 

4 . 3 Mixtures of Gases. This identification of the distri- 
bution-constant with the inverse temperature is further 
confirmed by studying statistically the distribution in phase 
of molecules of different kinds in a vessel. 

Let us deal with a mixture of two gases, n molecules of one 
and n' molecules of the other. Consider a state in which 
n l9 n 2 , ...... , n c molecules of the first gas are in the c 

phase-cells ; n\, n' 2 , ...... , ri c of the second also in the 

same phase-cells. Now such a state can be produced by 
associating any complexion of the (n l9 n 2 , ...... , n c ) state 

of the first gas with any complexion of the (n' lt n' 2 , ...... , 

ri c ) state of the second. Thus the number of complexions 
embraced by the (n lf n 2 , ...... , n c , n' v ri 2 , ..... , ri c ) 

state of the mixture is the product of the separate com- 
plexion-numbers for each gas. That is 

W (n lt ?i 2 , ...... n e9 n' l9 n' 2 , ...... n' e ) = 

n\ ___ ri\ 
n \ n 2 l ...... n c \ n\l 



50 STATISTICAL MECHANICS FOR STUDENTS 

and by the Stirling approximation 

log W = n log n + n' log n' J Sn T log n, + Sn' t log n' r j 

(4.3.1) 

To find the normal state we must determine the values of 
the n r and ri r , giving a maximum value to log W subject to 
the conditions 

Sn r =n . . . . (4.3.2) 

Zn' r = n' . . . . (4.3.3) 

and 27<r r n r + ZV r n' T = E . . . . (4.3.4) 

where E is the energy of the mixture. Careful attention 
should be paid to the form of condition (4.3.4). We have 
not two energy conditions 

27e r n r = constant 
2V r n' r = constant 

for the energy of any one component of the mixture does not 
remain constant. The intennolccular collisions involve a 
perpetual exchange of energy between the molecules irre- 
spective of the group to which they belong, and so the energy 
condition is expressed in one equation not two. It is this 
feature of the equations which is responsible for the important 
result which we shall deduce presently. (There is in general 
no equality between the energy of one type of molecule in 
a given phase and that of the other type in the same phase 
since 

2 I ~ 2 I (2 
r ' r r . 

r ~ 2m 

and , _ r 2 + i)* + C 2 

r 2^ 

quite apart from considerations of potential energy.) 

On proceeding to- the solution of the problem on the same 
lines as before, we find we have to satisfy 

2(1+ log n r ) 8n r + Z (1 + log n' r ) 8< - 
8n r =0 
2Bn/ - 
E T 8n r + 2V Src/ = 



TEMPERATURE AND DISTRIBUTION CONSTANT 51 

Using the method of undetermined multipliers we multiply 
the second equation by A, the third by A', and the fourth by 
//, and add. The 2c + 3 quantities, viz., the " normal " 
values of the n r and n r ' and A, A', p are determined from 
(4.3.2), (4.3.3), (4.3. 4) and the 2c equations 

log n r + A + fie, = . . (4.3.5) 
log n,' + A' + /**/ - . . (4.3.6) 

(r = 1, 2, ...... c consecutively). 

It is to be noted that, for the reason mentioned, the same 
fji occurs in the c equations (4.3. 5), as in (4 . 3 . 6), although 
not the same A. These equations yield the normal values, 
n r = v r and n/ = v r ' where 



C being e~* and C', e~ v ; C, C' and //, are, of course, 
worked out in detail in terms of E, n, ri , and the parameters 
involved in the r arid / by means of (4.3.2), (4.3. 3), 
(4.3.4). 

Thus the presence of one gas docs not upset the nature of 
the normal distribution of the other, and, as a significant 
fact, the same value of distribution-constant appears in the 
normal state of each part. The methods of sections (4.1) 
and (4 . 2) are once more applicable, and we find that the 
average energy of any molecule irrespective of type is as 
before, 1-5 /z," 1 and the temperature is (k^)" 1 . 

If the two molecular systems were in separate enclosures 
they would for an extremely large part of their history be 
distributed in or near the normal state, but not necessarily 
with the same distribution-constant. However, on mixing 
them, their normal state now involves a common-distribu- 
tion-constant. This statistical deduction, which as we have 
seen is derived from the dynamical principle of the con- 
servation of energy, is the analogue of the attainment of a 
common temperature by two gases on mixing. The statis- 
tical result can clearly be extended to a gaseous mixture of 
any number of different gases. 

In this connection Avogadro's hypothesis can be readily 
deduced from these results. Since the two gases at the same 

E2 



52 STATISTICAL MECHANICS FOR STUDENTS 

temperature have the same distribution-constant for their 
normal states, it appears from (4.2.3) that they have the 
same k. Therefore, the gas constants, R, for two quantities 
of each gas, each quantity containing the same number of 
molecules, have equal values. But this amounts to saying 
that equal volumes of two gases at the same temperature 
and pressure contain equal numbers of molecules. For a 
gram -molecule of any gas, R is known to have the value 
8 32 x 10 7 ; and the number of molecules in a gram- 
molecule is known to be 6-06 X 10 23 . Hence the value of k 
the gas-constant per molecule, is 1-37 X 10~ 16 . 

From (4.2.4) 



and so the specific heat of a monatomic gas at constant 
volume is 3 R/2. 

4 . 4. Potential Energy. If an external field of force (other 
than the one postulated to limit the size of the gas) is acting 
on the molecules, the integrations carried out in section 
(4.1) require some little modification, since e is now the sum 
of (f 2 + f] 2 + 2 )/2m and a function $ (x, y, z), and the 
equations (4.1.3) and (4.1.4) become 



y = n 

^ (4.4.1) 

D[ ..... f <f> e-^dx dy dtdt d, dC 



exp 
cZefydC = E . . . (4.4.2) 

The first integral in (4 . 4 . 2) is the potential energy of the 
whole gas in the external field of force. Call this <i>. Also 
represent the integral 



me 



cfe 



throughout the whole volume of the gas by the symbol F. 
Then in (4 . 1 . 5) and (4 . 1 . 6) we replace D by DF and E 
by E 4> ; so (4 . 1 . 7) is replaced by 



TEMPERATURE AND DISTRIBUTION CONSTANT 53 



2 (E - <&)' 

so that /i" 1 is still equal to two-thirds of the average kinetic 
energy of a molecule and the relation between the tempera- 
ture and the average kinetic energy is not disturbed. The 
existence of the field does not affect the distribution of the 
velocities among the molecules ; it does, however, produce 
lack of uniformity in the density of molecular concentration, 
since the density in a small volume surrounding the point 
(x, y, z) is proportional to 



and thus decreases as we move to places of higher potential. 
The reader should rightly appreciate these statements. In 
a region at high potential there are not, of course, in the 
normal state, as many molecules within certain limits of 
velocity as in an equal sized region at a low potential ; but 
the ratio of the numbers of molecules within two defined 
extensions-in-velocity in a given volume is unaltered by the 
field of force. 

4 . 5 Some Remarks on Pressure. The method of calcu- 
lating the pressure in section (4 . 2) is probably familiar to 
the reader, who has no doubt already met it in some text- 
book of physics. Nevertheless some further discussion of it 
may not be out of place here, although it might have pro- 
duced a rather long digression in section (4.2) from the main 
object of the chapter. 

It should be realised, of course, that the transference of 
normal momentum is not to be considered as only from one 
side of the element of area A to the other ; that would only 
have given half the pressure. Calling the two sides of the 
area a and 6, we have to estimate the components of 
momentum normal to the area towards 6 transferred from 
side a to side b per unit time, and the components towards a 
transferred from b to a. These two parts are of course equal 
in statistical equilibrium, and their sum is the pressure. 
The procedure can be related very easily to the usual defi- 
nition of pressure as force per unit area ; for if we conceive 
a physical surface situated in the interior of the gas, molecules 



54 STATISTICAL MECHANICS FOR STUDENTS 

moving up to the surface and those which have just re- 
bounded from it correspond to the two streams of molecules 
for an element of area parallel and near to the surface, and 
the algebraic change of momentum in a given time pro- 
duced by the surface is the arithmetical sum of the normal 
momentum -components towards the surface for the on- 
coming molecules and normal components away from the 
surface for the receding molecules. 

In view of similar considerations arising later in con- 
nection with liquids, it is very necessary to be quite clear in 
each particular illustration as regards the meaning to be 
attached to the word " pressure/' Besides the idea of rate 
of transference of normal momentum, there is the familiar 
notion of " hydrostatic pressure " as force per unit area on 
a surface arising from a " body " force, such as gravity 
being exerted on each element of the fluid. It should be 
obvious that so far we have not had much to do with that 
conception. In any case if it is a question of a relatively 
small quantity of enclosed gas, its weight is negligible in the 
treatment of its pressure, since it merely produces a small 
excess in the pressure exerted by the bombarding molecules 
on the lower part of the flask over that on the upper, because 
of the slightly larger density. Of course in the treatment of 
our atmosphere as a whole, the two ideas lead to the same 
value ; for it is the earth's gravitational attraction which 
holds our atmosphere to it and as the pressure (in the 
" bombardment sense ") vanishes at its outskirts, its pressure 
at lower levels is just equal to its hydrostatic pressure, 
estimated in the elementary way as weight exerted on unit 
area. 

It is, however, in the case of a liquid that serious con- 
fusion will arise unless care be taken. As we shall see later, 
the notion of an " internal " or " intrinsic " pressure due to 
transference of momentum across an element of area in the 
interior arises in theoretical discussion just as in the case of 
a gas ; but on account of the larger density this pressure is 
relatively enormous, and we have no direct experience of it. 
When we use the phrase " pressure of a liquid," we refer, as 
a rule, to the hydrostatic pressure arising from its weight 



TEMPERATURE AND DISTRIBUTION CONSTANT 55 

and varying markedly with depth in the liquid. But as a 
matter of fact we shall see later that in the statistical dis- 
cussion of the liquid state the symbol p is not even used to 
represent this magnitude either, but is rather associated 
with the pressure of the saturated vapour of the liquid. In 
due course we shall see why this is so, and also how it comes 
that the relatively great internal pressure is not capable of 
direct experimental detection. 

There is in the case of a gas an interesting and instructive 
way of connecting pressure regarded as rate of transference 
of normal momentum per unit area and pressure regarded 
as force per unit area. It will be recalled that, instead of 
considering a physical boundary to a quantity of gas, we 
imagined a field of force acting normally inward on the gas 
molecules at its external layer. We also saw in the last 
section that if </> (x, ?/, z) is the value of the potential energy 
of a molecule in this field at the point (x, y, z), then the 
density of the gas at this point is p where 

P=A,e-'* 

p being the uniform density of the gas throughout its 
volume with the exception of the exterior layer referred to. 
Now it is proved in text-books of hydromechanics that 
if p stands for the pressure of a fluid in the hydrostatic 
sense, then p is a function of position which is connected 
with the body-force by the following equations 

X=i?? 

p dx 

Y =!? 

pdy 

Z=l^ 

p dz 

where X, Y, Z are the components of the body force per unit 
mass of the fluid. Since X, Y, Z vanish in the interior of the 
gas, the hydrostatic pressure, p, is uniform throughout it. 
In the layer, however, 

X = - I 9 + etc., 

m ox _ 



56 STATISTICAL MECHANICS FOR STUDENTS 

m being the mass of a molecule, and so 

ty _ P ^ 
dx mdx 



= _, 

m ex 

Hence 



the constant of integration being zero since p vanishes as (f> 
approaches infinity. Hence in the interior of the liquid, 
where $ is zero, 

P =J- 

m p. 

or n 

pv = - 

M 

where n is the number of molecules in volume v. But this 
is just the same equation as (4 . 2 . 1) where p, however, 
represented pressure in the sense of rate of transference 
momentum. 

Equation (4.2.1) shows that if u 2 is the mean of the 
squared velocities of the molecules, then 

p = _pu 2 
^ 3^ 

since E = - 2 mu* =p vu*. 
2 2i 

Thus the " mean squared velocity " is equal to (3 #//>)*, 
and so is of the same order of magnitude as the velocity of 
sound through the gas ; for the latter is known to be 
(i<p/p)* where K is the ratio of the specific heat at constant 
pressure to that at constant volume, and is between 1 and 
1-66 in value. For hydrogen at C., the mean squared 
velocity is 1,840 metres per second, for oxygen 460 metres 
per second, etc., varying in fact inversely as the square 
root of the density under standard conditions or inversely 
as the square root of the molecular weight. 



CHAPTER V 

EXTENSION TO MORE COMPLEX MOLECULES 

5 . 1 Equipartition of Energy. Without the assistance of 
general dynamical theory, any treatment of molecules which 
are endowed with a structure must in the nature of things 
be inadequate. However, the author's aim being to intro- 
duce the reader to statistical results of general interest 
with as little delay as possible, leaving a more thorough 
treatment to a later stage, we shall make shift to consider 
the case of complex molecules with the aid of a simple 
model which has rendered yeoman service in the past 
to students not too well equipped with mathematical 
artillery. 

Let us regard each molecule as containing within itself 
" oscillators," i.e., particles attracted by elastic forces 
towards centres, where they would remain in relative 
equilibrium with respect to the general body of the molecule. 
The restoring force is supposed in each case to be pro- 
portional to the displacement of the oscillator from its 
centre in the molecule and so the oscillation is harmonic. 
Let q stand for the displacement ; the velocity of this dis- 
placement, dq/dt, is represented by q and the acceleration, 
d 2 q/dt 2 , by g. If a is the mass of the particle and bq the 
restoring force, then the equation of motion is 

aq + bq = 
and the solution of this is 

q = A sin (cot 6), 

where co is equal to (b/a)* and is called the "pulsance" 
(2 TT X frequency) of the motion. A is the amplitude of the 
vibration, and 6 is the " epoch-angle " determined by the 
fact that at times t = 0/eo, 0/co TT, 0/co 2 TT, etc. the 

67 



58 STATISTICAL MECHANICS FOR STUDENTS 

displacement is zero. The kinetic energy is aq 2 and the 

1 2 

potential energy is bq 2 . The total energy is therefore 

Z 

-b&? sin 2 (a>t - 0) + i a A 2 a> 2 cos 2 (arf - 0), 
2t 2i 

or a A 2 a> 2 . The momentum aq we shall represent by the 

Zi 

symbol p. 

The amplitude and the epoch-angle are not determined 
by the equation of motion ; they depend on the so-called 
initial conditions ; e.g., in the case of a simple pendulum, 
while the time of swing is determined by the length and 
intensity of gravity, the amplitude and the actual instant at 
which the string has an assigned inclination are arbitrary. 
A and 6 are, in fact, two arbitrary integration constants 
which enter the solution during the integration of the 
equation of motion. Physically this means, as regards the 
oscillator, that A and have definite values during a free 
path of the molecule which contains it, but are altered at 
every encounter between this molecule and any other. In 
any free path of a molecule the internal oscillator is engaged 
in a harmonic oscillation with a definite pulsance cu, but 
with an amplitude and epoch-angle which vari^s^g^Pfeee 
path to free path. To make this more obvious we may 
consider a molecule as a frame of reference for the oscillator. 
The set of axes with reference to which we estimate q are 
fixed in the molecule. (For the moment we are regarding 
the oscillator as having one " degree of freedom/' i.e., 
vibrating to and fro parallel to one axis). As long as the 
frame of reference is moving with a uniform velocity, the 
oscillation is not interfered with ; at time t 

q = A sin (a>t 6) 
q = Ao> cos (cut 6). 

An encounter produces an acceleration in the molecular 
frame of reference for a brief time. This is equivalent to a 
force acting on the oscillator for the same time in the 



EXTENSION TO MORE COMPLEX MOLECULES 59 

opposite direction. We are here appealing to the well known 
mechanical aspect of the relativity principle. We have only 
to consider our experiences in a carriage, which has been 
moving steadily and is then suddenly accelerated or retarded, 
in order to realise the situation. The oscillator suddenly 
receives an impulse in a direction opposite to that on the 
molecule. When it is over, the state of affairs is such that 
the oscillator has practically the same q but a different q 
and for the subsequent spell of harmonic motion before the 
next shock, q and q cannot be given by the equations above, 
but by two such as 

q = A' sin (a>t 6') 
q = coA' cos (cut 0') 

with a different A and 9, but, of course, with the same 
pulsance as before. Hence at every collision not only is the 
kinetic energy of the molecule changed, but also the internal 
energy of the oscillator also. The collisions, in short, effect 
an exchange of energy of translation between the molecules 
and also between this energy and the internal energies of the 
molecules. 

We can as before bring this new " degree of freedom " 
within our considerations of probability. The phase of a 
molecule is now determined by eight components, x, y, z, q> 
> ??> C, p, and an extension-in-phase by the limits x to x + 

8x, , q to q + S#, to + Sf , , p to p + Sp. 

The phase-diagram is now an eight-dimensional one ; or if 
it is preferred, it can be visually represented by four plane 
diagrams, a point in the fourth one representing q and p. 
A phase is represented by a " point " in the phase-diagram 
or a four-point group in the four plane diagrams. A phase- 
cell can be pictured if one likes, as a group of four elementary 
rectangles and a " path " as a group of four curves. 

If now one " aspect " of a molecule is as possible as any 
other with this extended notion of aspect, the probability 
of a state in which n^ representative points are in the first 
cell, etc., is as before given by the quotient of W (n l9 n%> 

, n c ) by the total number of complexions consistent 

with the energy condition. The procedure is just as before ; 



60 STATISTICAL MECHANICS FOR STUDENTS 

the state of maximum probability is given by equations 
(3 . 2 . 1), (3 . 2 . 2), (3 . 2 . 3), where e is defined to be 

2 + ^ + C 2 , P 2 , , , x , 1 L 9 

! - \ I - \ - -J- _ -j- J) (x ,y , z) + - fr? 2 , 
2m 2a y v J J 2 *' 

and involves the phase (<?, ^) and parameters (a, 6) of the 
internal motion. The same type of proof also serves to 
show that this state and those very near it constitute the 
normal state of the system of molecules. 

5 . 2 Partition of Energy. On reverting to the procedure 
of section (4 . 1), we can easily obtain average values for the 
various parts of the kinetic energy of a molecule which are, 
as we say, associated with its various degrees of freedom. 

In the first place the equation (4 . 1 . 3) is replaced by 

D[ ..... fe"^ Ax dy dz dq d drj d( dp = n . (5.2.1) 
In the absence of an external field this reduces to 

e~w* dp = n 



where the integrations are practically from GO to + oo 
in each case * and 

a=A)8=!/*&, y = -^. 
2m P 2 P ' r 2a 

We thus have 



The kinetic energy associated with the motion of a 
molecule parallel to the axis of x is 2 /2ra, and so the average 
of this over every molecule at any moment is 



~" d * ...... dp 

divided by n ; i.e., 

^ j> e-* 1 dt . Je-" f dr, ...... Je-^ 1 dp 

* See Note 1 to this chapter. 



EXTENSION TO MORE COMPLEX MOLECULES 61 

divided by the expression on the left-hand side of (5.2. 2). 
This is equal to 



2m 2 

1 

4 m a 
1 



A similar result follows for the average of ?? 2 /2ra and 

and also for the average of the part of the kinetic energy 

p 2 /2a associated with the internal movement. 

The reader should carefully note the meaning of the word 
" average " in this connection. No statement is made about 
the average total or partial energies of an individual mole- 
cule over a finite lapse of time. We cannot follow individual 
histories. The " equipartition law " is concerned with 
energies averaged over all the molecules at any one instant. 

The modification introduced into the treatment of pressure 
in section (4 . 2) owing to the considerations of internal 
structure is easily dealt with. It is not difficult to see that 
the pressure is now two-thirds of the density of the kinetic 
energy of translation of the molecules (not of the whole 
kinetic energy) ; so by the result just obtained 



3 

n 



71 

" r> 



2,* 



* 



3 v 

n 

or pv = - 

V* 
where p is the pressure. As before, this result leads to the 

* p is here the pressure and must not be confused with the momentum 
of an oscillator. 



62 STATISTICAL MECHANICS FOR STUDENTS 

identification of the distribution constant p with (k6)~ l , 
where k is the gas constant per molecule. 

Indeed the reader should have no difficulty in seeing that 
the line of proof used can be extended to molecules with 
oscillators having more than one degree of freedom, i.e., 
free to vibrate in all directions with reference to axes fixed 
in the molecule with three degrees. Furthermore, molecules 
containing more than one oscillator, each oscillator having 
its own distinctive mass and force-constant, can also be 
brought within the ambit of the proof. And lastly mixtures 
of different molecules can be treated in a similar manner to 
that used in section (4. 3).* The striking feature is the 
appearance of the same distribution constant in the exponen- 
tial factor of the distribution law, and this as we have seen, 
is the result of the liberty of exchange of energy between all 
the degrees of freedom of the molecules as a whole and of 
the internal oscillators. The equality between /x" 1 ad kO 
is still maintained and the average kinetic energy associated 
with any degree of freedom is as before | kO. If p, p' ', p" ', 

are the partial pressures of the constituent gases, 

then 

n 
pv -- nkv 

P 

p'v = n'kB 
etc. 

where n, ri, are the numbers of each type of mole- 
cule present, this being the familiar law of partial pressures. 
It is very necessary to observe that the equipartition law 
has been confined to average kinetic energies, and it is easy 
to see that the restriction is connected with the fact that the 
kinetic energy of each molecule is the sum of terms each of 
which depends on the square of a momentum, involving in the 
proof the possibility of splitting a certain part of the expres- 
sion e~** into separate factors of the type e~ a *\ , 

e"^. It is true that this is just as much a feature of the 
potential energy of the oscillator in a molecule, which is 
given by a term involving the square of a co-ordinate ; and 

* See Note 2 at the end of the chapter. 



EXTENSION TO MORE COMPLEX MOLECULES 63 

we can prove just as before that the average potential 
energy of a harmonic oscillator is also ^ kd for each degree 
of freedom. But in general potential energy is not given by 
simple square terms. If the restoring force on the oscillator 
were not proportional to q, the oscillator would be " anhar- 
monic," and the potential energy would not be represented 
by \ bq 2 and the equipartition proof would fail for it. Also, 
as regards the external field of force, the potential energy 
function <f> (x, y, z) is not in general a sum of squares and, 
again the proof would fail in this connection. For example, 
if it were a uniform field parallel to the axis of x, the potential 
energy would be proportional to the -co-ordinate (choosing 
the origin at a suitable level) and writing Ax for cf> (x, y, z) 
we would obtain the average potential energy by dividing 

DA f x e-^ Ax dx( (e~^' dy dz dq d dj] dC dp 

by . 

D l c"^^ I \e~^' dy dz dq d drj d dp 

Jo J J 

where e' = e Ax. 
This is 



r 



which is equal to 

2 



i.e., juT 1 or kd, just twice the average kinetic energy. In fact, 
if the potential energy were proportional to any power of 
x, say x n , then it can be easily deduced that the average 
potential energy would be k6/n. As a general rule, even 
such relatively simple expressions as sums of powers of 
the co-ordinates do not hold sway, and so the reader must 
be careful in the use of the equipartition law, unless applied 
to kinetic energy. Later, when we reach a fuller dynamical 
treatment, we shall see that there is a general partition law 
which covers the equipartition of kinetic energy and gives 



64 STATISTICAL MECHANICS FOR STUDENTS 

us some idea as to procedure in other eases. All this is, of 
course, based on classical dynamical laws. Wider con- 
siderations involving the quantum hypothesis will modify 
even this general law. 

We cannot leave the subject of complex molecules even 
at this stage without some reference to molecules involving 
more than one atom. A familiar picture for certain diatomic 
molecules is that of a dumbbell two separate atoms held 
rigidly together. This picture introduces new degrees of 
freedom connected with rotation as distinct from trans- 
lation. The student may naturally remark that rotation is 
just as much a possibility for monatomic molecules as for a 
diatomic and should therefore have been considered earlier. 
The answer to this is to draw the reader's attention to an 
implicit assumption in the model of a monatomic molecule 
used hitherto. It has been regarded as dynamically equi- 
valent to a hard smooth sphere. The mutual forces exerted 
at encounters are normal to the surface, and passing through 
the centre produce no change in rotational momentimi nor 
in the individual rotational energies of the molecules. Thus 
rotational energy is not exchanged between molecule and 
molecule at an encounter, and since it therefore makes its 
appearance in the equation (3 . 1 . 2) as a constant, it docs not 
appear at all in (3 . 2 . 8). But it is otherwise for a dumbbell- 
shaped particle ; the forces between the atoms of different 
diatomic molecules will not produce any change in the com- 
ponent of rotational momentum around the axis of figure, 
but will do so about any axis at right angle to the figure- 
axis. This amounts to excluding one of the three new degrees 
of freedom.* Again we must anticipate later dynamical 
work when we state that each effective degree brings in on 
the average ^ k9 of kinetic energy of rotation. Of course, for 
a less symmetrical diatomic molecule or for a polyatomic 
molecule, provided we can regard it as rigid, we should have 
to take into account three degrees of freedom for rotation. 
These remarks have an important bearing on the specific 

* The reader is reminded that rotational velocity about any axis can 
be resolved into three component angular velocities about threo Cartesian 
axes of reference. 



EXTENSION TO MORE COMPLEX MOLECULES 65 

heat of a gas. If the gas be monatomic, with no internal 
degrees of freedom, the energy -content will be 1*5 nkff, and 
so the heat capacity at constant volume will be 3R/2 where R 
is the gas-constant for the amount of gas considered. This 
is in good agreement with facts. For a diatomic gas we 
would expect kO more energy on the average in a molecule, 
and this would lead to 5R/2 as the heat-capacity. For more 
complex rigid molecules we would expect a still further \ k0 
of energy per molecule involving 3R as the value of the heat- 
capacity. Provided temperatures are not too low or too 
high, several gases show good agreement with these results. 
Thus from elementary thermodynamical reasoning we know 
that the heat-capacity of a quantity of gas at constant 
pressure is greater by R than the heat-capacity at constant 
volume, and so the ratio of the two specific heats for diatomic 
gases should be 7/5 or 1-4, and for more complex gases 4/3 
or 1*33, two results which are in good agreement with the 
facts for some gases. But as diatomic gases are reduced in 
temperature, it is found that their specific heat per gram- 
molecule falls asymptotically to the values for monatomic 
gases, from which it would appear t* t for some reason not 
evident in the classical dynamical i , vtment, the rotational 
energy has on the average a value p jressively smaller and 
smaller than the amount kd per molecule. This is a dis- 
crepancy which, as we shall see later, involves the use of the 
quantum hypothesis to remove. At high temperatures the 
specific heats show signs of attaining higher values than 
those theoretically deduced. To account for this we have 
to abandon the restriction as to rigidity in the molecule and 
admit that the molecule may have kinetic energy arising 
from the relative vibrations of its constituent atoms and 
potential energy arising from relative displacements. But 
again we have the same difficulty to meet. The classical 
treatment does not show why these energies of internal 
vibration and strain should not make their appearance at 
normal temperatures. So long as elastic strain of the parts 
of the molecule are postulated the relative partition of the 
total energy among the various degrees of freedom is settled 
not by the particular value of p, but by the expression for 



66 STATISTICAL MECHANICS FOR STUDENTS 

the energy of a molecule in terms of its co-ordinates and 
momenta (external and internal). Thus for all " squared 
terms " there is partition on an equality basis ; and for 
other terms though not so simple, it is definite. Nor is this 
all. The facts of the spectral lines of gases show that there 
is a quite complicated vibrational mechanism within every 
atom and molecule which is the dynamical equivalent of 
numerous oscillators. Why then should the specific heat of 
any gas not be very much greater than it really is ? On 
any reasonable assumption in addition to the half-dozen or 
so energy-quantities of |- TcO assigned to each molecule, there 
should be many more for the internal degrees of freedom of 
the radiating mechanism. Again these discrepancies can 
only be dealt with by an appeal to some form of quantum 
theory and there we must leave it for the present. 



NOTE 1. Since the oscillator must in its vibration be 
confined within the molecule, it may appear absurd to 
integrate for q from oo to + oo. Actually it is not incon- 
sistent with our knowledge of the sizes of atoms and molecules 
to conceive that a displacement of the oscillator is quite 
possible which renders the potential energy very much 
greater than the average value (2 j^)" 1 ; for such a displace- 
ment e~ M * would have reached such a minute value that 
the part of the integral between this value #nd infinity is 
negligible. A similar remark applies to the p integration. 

NOTE 2. A slight difficulty in the mathematics may 
present itself to the reader in connection with the fact that 
the number of oscillators or degrees of freedom within a 
molecule of one type in the mixture may not be the same as 
the number within a molecule of another type ; so that we 
cannot apparently use a common set of phase-cells for each 
group of molecules, the dimensionality being different in 



EXTENSION TO MORE COMPLEX MOLECULES 67 

each case. Formally we surmount the difficulty easily by 
bringing the number of oscillators in every molecule up to 
the same value. We can then conceive the " extra " oscil- 
lators in those molecules which have really less than the 
maximum number to possess zero mass and so to contribute 
no energy to the total amount. 



?2 



CHAPTER VI 

THE SECOND LAW OF THERMODYNAMICS 

6 . 1 The Normal State of a Molecular System and Thermo- 
dynamic Equilibrium. In accordance with our aim of 
bringing out the connection between statistical mechanics 
and thermodynamical facts as soon as possible, we propose 
in this chapter to deduce the second law of thermodynamics 
for a gas from the results hitherto obtained, before proceed- 
ing to a further development of the subject itself. 

We have seen that the molecular system is, if our prob- 
ability-postulate be accepted, for a relatively great part of 
its history in or extremely near to the " normal " state 
defined by the c equations 

Vf = C e"*"' 
or 

log V r = A fJL r (6.1.1) 

where A is written for log C. (For convenience a change of 
sign in A from the earlier sections is made.) These, combined 
with 



v=l 

c 



. . . . (6.1.2) 



Z r v r = E (6.1.3) 

serve to determine, the v r > A and //, as functions of E, n and 
the parameters which we shall denote by the symbols 
a l9 2 , . . . . , a e . Let us change our point of view a little and 
regard (6.1.1), (6.1.2) and (6 . 1 . 3) as equations deter- 
mining the v r , E and A as functions of /z and the parameters. 
That is, we are going to consider the molecular system 
passing from the normal state for one value of ^ to the normal 



68 



THE SECOND LAW OF THERMODYNAMICS 69 

state for another value of ft, involving, of course, changes in 
the energy E, as well as in A and the individual v r or normal 
numbers in each phase-cell. Remember that p has no 
meaning for the molecular system apart from the normal 
state. If the system be not in the normal state, it is a direct 
result of our postulate that it will gradually tend to it. There 
is no dynamical impossibility involved in a statement that 
there might be abnormal states of distribution in which the 
system might remain for ever. All we can say is that on our 
probability basis it is enormously improbable. It is not 
impossible that all the peoples of the world could be on the 
Isle of Man in one day (actually a density of one person per 
square yard would just about suffice) but well ! the point 
need not be laboured. 

The analogy between this normal state and the state of 
thermodynamic equilibrium of an isolated physical system 
is too obvious to escape notice, and the analogy is very close 
indeed ; for from the equations of the normal state we can 
construct functions of /z and the parameters which are as 
a matter of pure mathematics connected by differential 
equations of precisely the same form as those which are found 
by experiment to connect those functions of the thermo- 
dynamic variables of a system which we call the internal- 
energy-function, the free-energy-function, and the entropy- 
function. The proof which follows is, to be sure, limited at 
the moment to gaseous systems ; but it is surprising how 
little further elaboration of mathematical detail is required 
when we deal with this question for other systems at a later 
stage. 

6 . 2 The Entropy Law. On solving (6 . 1 . 1) (6 . 1 . 2) (6 . 1 . 3) 
we express the v r as functions of ft, a l9 a 2 , a e (remem- 
bering that l e 2 , , c are functions of a v a 2 , , 

aj. Inserting these in (6 . 1 . 3) we obtain a function of ft, a x , 
2 > , # e > which is equal to E ; we shall denote it by 

H (/x, a v a 2 , , a e ) 

or briefly by 

H (p, a). 

Also from (6.1.1) we express A as a function of the ft, and the 



70 STATISTICAL MECHANICS FOR STUDENTS 

a r . Let us write for n A//H. the functional form ^P (/*, a v a 2 , 
...... , a e ), or 



Suppose the molecular system to change from the normal 
state for the distribution constant JJL and parameters a v a 2 , 
...... , a e (the statistical-mechanical (S.M.) variables) to 

the normal state for values /U, + 8/i, a l + Ba l9 a 2 + 8a 8 , 
...... , a e + 8a e of the S.M. variables. The energy will 

alter to a value E + SE where 

8E = H (p + fy, a + 80) H (/*, a). 

Of this change in the energy, a certain part is given by 



c l oa 2 

We shall denote it by SE l9 so that 



. . (6.2.1) 



This part arises from the change in the parameters, but 
with the distribution still kept at the original normal state 
(n r = v r ) and not altered to the new normal state (n r 
v r + 8v f ). The remainder of the energy 8E 2 ( = SE SE X ) 
is given by 

8E 2 = i;J:6 r ^8a, + i:c r ^s M . . . (6.2.2)* 

r=l t i ca 9 r -l CIJLC 

The first part, SE l9 does not arise from effective changes 
in the co-ordinates and momenta of the molecules ; for the 
numerical distribution of phases among the cells is regarded 
as unaltered in estimating SE 1 . The part 8E 2 does arise 
from the changes in cell-distribution occasioned by the 
changes in the parameters and ^. It is plausible to regard 
this second part as the analogue of the heat supplied to the 
system and the first part as the analogue of energy trans- 
ferred to the system by purely mechanical means. We recall 
the fact that among the parameters are a group related to 

* Remember that the ( r do not depend on ,u. 



THE SECOND LAW OF THERMODYNAMICS 71 

an external field of force which limits the volume of the 
molecular system ; any change in these involves alteration 
in the boundary and work of an " external pressure." Thus 
SEj may be regarded as the analogue of " external work " 
done on the system. 

Quite apart from such interpretation, however, it can be 
deduced as a matter of mathematics only that, with SEj 
and SE 2 defined as in (6 .1. 1) and (6 .1, 2), 

/.SE 2 = 8{^[H(^a)-^( M ,a)]} . . . (6.2.3) 
We shall defer the actual mathematical steps for a moment 
so as not to interrupt the general line of thought. Suppose 
the molecular system to experience a finite change from a 
normal state (p,', a') to a normal state (//,", a") passing through 
all the intermediate normal states on its way (just as in thermo- 
dynamic reasoning a physical system is supposed to pass 
from state to state through intermediate states of thermo- 
dynamic equilibrium), then 



= //[H (//, a")- V (,/, a")] - / [H (//, a') - ^ (//,a')]. 

Thus the integral of JJL d E 2 from the state s' to state s" along 
a track of normal states depends only on the initial and final 
states. The analogy with the entropy-theorem of thermo- 
dynamics is obvious. On interpreting 8E 2 as heat supplied 
to the system, and (kfji)~ l as the temperature 0, we have 



> (ju/, a') 

J 9 > U 

where 

and is the analogue of the entropy -function of the thermo- 
dynamic state of the physical system. 

To complete the analogy we have to discover the statis- 
tical-mechanical theorem corresponding to the increase of 
entropy which takes place when an isolated physical system 
passes in an irreversible manner from one state of equilibrium 
to another. Such an irreversible change takes place through 
intermediate states some of which at least are not states of 



72 STATISTICAL MECHANICS FOR STUDENTS 

equilibrium. Hence in the statistical-mechanical analogue 
we must conceive of some way in which a molecular system 
in a normal state may pass to another normal state through 
intermediate states which are not all normal. This can easily 
be effected. In the thermodynamic processes we alter the 
thermodynamic variables from the values which hold for the 
first state to those which hold for the second in a short time. 
This ensures irreversibility of the path. The essence of 
reversibility is the infinitely slow change in the variables 
allowing the system to pass through intermediate states of 
equilibrium. Similarly we quickly change the S.M. variables 
from the values //, a' for the first state to the values /z", a" 
for the second state. To carry through the reasoning we 
require to know another mathematical result, the proof of 
which we shall also defer for a moment ; it is this : 

3> (p, a) = k (log W m - n log n) . . (6.2.5) 

where W m is the maximum value of the function W (n l9 n^ 
n c ) being equal in fact to W fa, v 2 , v c ). 

Suppose then the molecular system is in the normal state 

characterised by (//, a\, a' 2 , , a' e ) ; a sudden change 

is made in the S.M. variables to the values //, a'\, a" 2 , 

. . . .v. . , a" e .* The distribution v/, i> 2 ', v c ' is no 

longer the normal distribution for the new variables ; the 
normal distribution is i//', j/ 2 ", , i/' c . 

Hence 

W fa', iV, v/)< W fa', /, c ") 

for the latter is the maximum value of the function W (n v n 2 , 

n c ) for the new parameters and distribution-constant. 

At the beginning of the change the system although in a 
normal state for the old S.M. variable is in an abnormal state 
for the new S.M. variables, and by our probability-postulate 
it will ultimately arrive at and keep very near to the normal 
state for these new values. It may experience all sorts of 
changes on the way ; W may fluctuate about in the most 
fortuitous way (we cannot follow the individual history of 

* Remember that there is no change of energy. The system is to be 
isolated. Hence changing parameters will in general involve a change of 
M also, so as to satisfy H (/*', a') = H (M*, *) 



THE SECOND LAW OF THERMODYNAMICS 73 

each particle), but presently the system will reach the new 
normal state with the increased value of W. If W m ' is the 
value of W (v 1 / , v 2 ', ...... , v e ') and W m " is the value of 

W (y/, v \ ...... , v c ") 9 W m ' was the maximum value for 

the S.M. variables in first state, just as W m " is for the 
second ; and since W m ' < W m ", therefore by (6.2.5) 



Thus in the statistical-mechanical argument the function 
4> (/x, a) behaves in a manner similar to the mode of behaviour 
of entropy in Thermodynamics. 

We have to fill in the missing mathematical steps leading 
to (6.2.3) and (6.2.5). They are three in number. 

I. Since any one of the equations (6 . 1 . 1) is an identity 
when the v r , X and the e r are regarded as functions of /* and 
the a,, the result of partial differentiation with respect to /z 
or any of the a, still yields identities. Hence we have the c 
equations 

3 log v r __ 3A ___ 



for any given a,. 
This is the same as 



__ ___ 
da s Sa 8 



dv r __ d\ _ 3e r 
da t r da 8 r da s 
On^adding the c equations we obtain 

dn dX c) r 

~n~ ^S v r ~^. 
ca 8 ca s r =i va g 

But, since n does not change, dn/da, = 0, and so 

Sv 8g r^ 8 ^(^ q ), 
r i f da g da s 

Thus 

Z v ***a. = S* lf <>'.*} Sa 

,_i ,=! da, ,,i da, 

But 

S* 0*. a) - -^ V 4, j?^0*.) 8a 



74 STATISTICAL MECHANICS FOR STUDENTS 
Hence 

SEi = 8V OK, a) ^P-^-' 8/i. 
As 8E 2 = SE - SE X it follows that 

SE 2 = SH (u, a) S^ (u, a) + ^-^ 8/t (6.2.6) 

3/A 

II. Reverting to (6.1.1) once more, we obtain the c 
equations 

9 log v r 3A 

5 == 5~ r> 

C//X. C7/X, 

or 3v r _ 3A 

which on addition yield the result 



So that 



CfJ, 

or 



_ _ (6.2.7) 



a/* ^ 

Combining (6.2.6) and (6.2. 7), we have the result 
8E 2 = 8[H 0., a) - * (M, ] + H ^a)-^0*.a) 



or ^ 8E 2 = 8 { /i [H (^ a) - ^ (/i, a] j , 

which is equation (6.3. 3). 

III. Lastly, on putting the values for v r in the expression 
for log W, we obtain 

c 

log W m = ft log n 27 v f (A p r ) 
ri 

= n log 7i /&A + /z E. 
So 

/i [H (/i, a) ^ (/i, a)] = log W m - n log w, 



THE SECOND LAW OF THERMODYNAMICS 75 

and 

<E) fa, a) = k (log W m n log n), 

which is equation (6.2.5). 

The molecular models which we employ thus give us a 
somewhat broader view of the property which we call 
" entropy " than pure Thermodynamics. The function 

W (n l9 n a , , n c ), is one which does not maintain a 

constant value in the actual history of a molecular system ; 
there are continual fluctuations in its value going on as the 

numbers n l9 n 2 , , n c change with the individual 

movements and encounters of the molecules. The probable 
amounts of these fluctuations we shall estimate later, but 
they are on the average small, although there is no dyna- 
mical impossibility in brief excursions to values very far 
removed from the normal v f . So if we adopt the terminology 
of Darwin and Fowler, and call 

A | log W (n l9 n 2 , , n c ) n log nj , 

the " kinetic entropy " of the system, we know that this 
kinetic entropy is fluctuating between its maximum value 
and values slightly below it, except for very occasional wider 
fluctuations to values well below. In thermodynamics, 
which is built on " macroscopic " observation, not " micro- 
scopic " analysis, we rest our reasoning on experimental 
results in which these fluctuations become " smoothed out," 
and an appearance of statical rest rather than of statistical 
equilibrium is presented, leading to the conception of a 
non-fluctuating thermodynamic entropy for whose value we 
naturally take the maximum value of the kinetic entropy. 
In thermodynamics, moreover, entropy is a property which 
can only be defined for states of equilibrium. In Statistical 
Mechanics, kinetic entropy is not so restricted to normal 
states alone. 

To conclude this chapter we have to point out that the 
function (/*, a) is the analogue of the free-energy function 
(at constant volume) of a thermodynamic system. Thus 
(6.2.4) can be written 

9 =H - 0*, 



76 STATISTICAL MECHANICS FOR STUDENTS 
and, moreover, 



dp, 



Id0 
/ dp 



which, by (6 . 2 . 7) = fyt (H ^) 

= 4>. 

These two results connecting W (/x, a) with H (//,, a) and 
4> (/x, a), are obviously formally similar to the mathematical 
equations connecting the free-energy function with the 
energy and entropy functions in Thermodynamics. 



CHAPTER VII 

THE ENTBOPY OF A PERFECT GAS 

7 . 1 The Problem of the Magnitude of a Phase-Cell. In 
Chapter IV. we carried through some mathematical opera- 
tions which involved the substitution of integrations of con- 
tinuous functions of phase for summations of a number of 
terms of a series. The nature of the problems which we were 
discussing at that point is such that no ambiguity results 
from this procedure. It is quite otherwise with the investiga- 
tion on which we are now about to fasten attention. In the 
immediately preceding sections we have obtained an expres- 
sion for the entropy of a monatomic gas ; it is 

* (log w m n log n) 



c 



or k 2 v r log v r . 

r=i 

Naturally we are concerned to discover the connection of 
this expression with the well-known thermodynamic expres- 
sion for the entropy of a gas, viz., 

s p log R log p + constant 

where s p is the heat-capacity at constant pressure. The two 
mathematical expressions do not bear a very obvious resem- 
blance to one another ; yet the conversion of the former into 
an integral leads quite directly after a few steps to a demon- 
stration of their essential agreement. But there is one step 
in the conversion which has been the origin of one of the 
most famous scientific discussions of the past two decades, 
and has led to results which are in some quarters regarded 
as one of the prime achievements of the quantum theory 
outside its triumphs in settling questions of atomic structure. 
In Chapter IV. we replace n r by a sextuple integral 



f f 



77 



78 STATISTICAL MECHANICS FOR STUDENTS 

where da is written as a symbol for the infinitesimal element 
of phase-extension 

d dr\ dCdx dy dz, 

and the integration is extended over limits determined by 
the r th phase -cell. Now we clearly cannot replace log n r by 

J ...... J log/efo, 

nor n r log n r by j ...... J / log / da. 

Strictly n r log n r is the product of I ..... I / da and 

log ...... / da, both integrations Extending over a 

phase-cell. If we wish to replace n* r log n r by an integral 
over the r th phase-cell, and not by a product of an integral 
and its logarithm, we must begin by introducing a symbol 
for the magnitude of the cell ; let it be g. Then n r /g is the 
average value of the function/ over the phase-cell ; it is the 
average density of the representative points in the cell. It 
follows that there is an approximate equality between log 
(n' r /g) and the value of log / at any phase of the cell, and so 
there is an approximate equality between the integral 



and the expression 



or n r log n f n r log g. 

It follows that there is an approximate equality between 

c 

Z n r log n r , 
and the expression 

J J /log/ da + nlogg, 

the integration being between the extreme limits of the 
phase-diagram. For the normal state we have seen in 



THE ENTROPY OF A PERFECT GAS 79 

Chapter IV. that the form of the function / (ignoring any 
external field of force) is 

Dexp {-a^+^ + f 2 )} 
where 

a= -^ . ..... (7.1.1) 

4mE v ' 

, T, n / 3n \ 3 / i \ 

and D = - ( - _ I . . . . (7.1.2) 

v v 



and so there is an approximate equality between 

c 

E v r log v r 

r=l 

and 



X flog D - a (* + ij 8 +P)ldv+nlogg. . (7.1.3) 

the integration extending between the widest limits of the 
phase-diagram. 

The quantity g is very vague at the moment. We only 
know two things about it. Although not a mathematical 
infinitesimal, it is " physically small " from the point of view 
of the experimentalist ; yet it must be large enough to 
contain very many representative points in the region of 
the phase-diagram favoured by the points in the normal 
state. Further its physical dimensions in terms of the 
fundamental quantities length, time and mass are the same 
as the cube of the quantity " action," which is defined as 
the product of energy and time, and is a concept of great 
importance in the general mathematical formulation of 
classical dynamical principles ; for the product of energy 
and time has clearly the same dimensions as the product of 
momentum and length. Nothing more can be elicited about 
g from classical sources. 

It is more than twenty years since Planck made a very 
definite proposal about its magnitude. He suggested that 
there is a precise definite value for g which makes the equality 
mentioned above not merely approximate but exact. Natu- 



80 STATISTICAL MECHANICS FOR STUDENTS 

rally such a statement could only be advanced with the 
support of some physical facts hitherto unnoticed. The 
facts which Planck appealed to were just those newly- 
observed phenomena which were at the time leading to the 
acceptance of a rather indefinite form of quantum hypo- 
thesis with its then revolutionary idea of a kind of atomicity 
in atomic occurrences previously unsuspected. There was 
also a new development in purely thermodynamic theory 
which was proving a fertile instrument of progress in the 
hands of Nernst and his pupils. Naturally we cannot enter 
into a discussion of these matters here, and must leave until 
later chapters some account of the repercussions on statistical 
mechanics of the quantum theory. But these remarks and 
the mathematical analysis in the next page or so will serve 
to prepare the reader for a fuller statement at a later stage. 
He is already probably aware that one fundamental feature 
of the quantum theory is the significance of a certain unit, 
or, if you like, " atom," of action known as " Planck's 
constant," and denoted by the symbol h. In terms of the 
erg-second as a working unit of action, the value of h has 
been determined in several ways to be 6'55 X 10~ 27 ; and 
to cut the matter short at the moment, the upshot of a great 
deal of discussion and experiment has been the suggestion 
that g = A 3 , since, as pointed out above, g has the dimensions 
of action cubed. Without taking any other step for the 
present than that of using g to represent a precise physical 
magnitude, let us proceed. 

7. 2 The Entropy Constant. By (7.1.3) the entropy of the 
gas becomes 



2 + ? + C 2 ) exp{ - a (f 2 + T? 2 + 
- AD log Df ...... f ea?p{- a (P + *? 2 + C 2 )}*r knlogg 



The first term is simply 2 JfcamE, or, by (7 . 1 . 1), 3 nk/2. 
The second term similarly reduces to n k log D. Thus the 
entropy of the gas becomes 

n k log D n k log g. 
2 



THE ENTROPY OF A PERFECT GAS 81 

On inserting the value of D from (7.1. 2), and completing 
a few obvious steps, we arrive at 

7(3, -nil 5, i 3 /, . , 4 77 m\ , ) 

n k - log E + log v - log n + - ( I + log ) log g[ 

{Z * A \ O / J 

Since E = 3 n k 0/2, we easily obtain for the entropy of 
a monatomic gas containing n molecules the expression 

7 ( 3 7 4 , , , , , (2 TT mkf . 3 ) 
ft k j - log + log v log n + log + - [ > 

(2 g 2) 

which, on changing from the variable v to p by means of 

^w = R0 = nkO, 

or log v = log log p + log ft- + log i 

becomes 

(5 (2 TT rafc) & 3 ) 

R \ log log p + log 1 > (7.2.1) 

(2 2j 

The formal similarity of (7 . 2 . 1) with the thermodynamic 
expression for the entropy of a gas is apparent. In view of 
Planck's hypothesis concerning the definiteness of the 
magnitude of g, it goes further than pure thermodynamics, 
for it suggests a precise value for the undetermined constant 
of integration which enters in the thermodynamic analysis. 
Now this is tantamount to asserting that a quantity of gas 
has at a definite pressure and temperature an absolute 
entropy, a position quite untenable in the classical thermo- 
dynamics of the nineteenth century in which only differences 
of entropy between two assigned states were considered. 
But, as mentioned already, Nernst's heat theorem, and the 
work carried out in his laboratory on measuring the affinities 
of chemical reactions and the specific heats of gases and con- 
densed materials at low temperatures, had in the early years 
of this century carried physical chemists at all events well 
away from the vagueness of the older position, even before 
the full blast of the quantum theory had played havoc with 
classical methods and conceptions. We cannot pursue this 
particular matter further at the moment. The main problem 
of this chapter has been to show the essential agreement 
between two formally very different expressions, and that 
has been achieved . We must pass on to further developments 
of statistical theory along classical lines. 



CHAPTER VIII 

THE STATISTICAL THEORY OF CHEMICAL EQUILIBRIUM 
IN A GAS REACTION 

8 . 1 Reactions Equivalent to a Simple Dissociation. Let 

us consider a system in which there exist atoms of two 
different types and diatomic molecules, each containing one 
atom of each type. The phase diagram represents the 
Cartesian co-ordinates of the centres of gravity of the 
dissociated atoms and of the molecules and the corresponding 
momenta ; it is, of course, six dimensional. Let there be 
present in all a atoms of one type and /? of the other ; so 
that if the dissociated atoms at any moment are a and b of 
each kind respectively, then the number of molecules at that 
moment is I, where 

a -f I =a /ft i n 

b + l=p ..... (8.1.1) 

It is assumed* that the formation of a molecule or its 
dissociation is accompanied by the liberation or absorption 
of a definite amount of energy. Thus, if there is no external 
field of force, the energies of the atoms and the molecules in 
a certain phase are given by 



2m 



. 



2 ! ^2 I 



__ 

2(m a + m b ) 

where w is a definite energy of dissociation and is regarded 
as positive if energy is absorbed in the dissociation of a 
molecule. 
In counting complexions we have to analyse the situation 



CHEMICAL EQUILIBRIUM IN A GAS REACTION 83 

more exhaustively than in earlier chapters. Let us consider 
the state in which 

#! atoms of first type, &J atoms of second type, 

/! molecules are in cell 1. 
a 2 atoms of first type, 6 2 atoms of second type, 

1 2 molecules are in cell 2. 



a c atoms of first type, b c atoms of second type, 

l c molecules are in cell c. 

The first step begins by supposing that we have a par- 
ticular atoms of the first type, b particular atoms of the second 
type, and I particular molecules to deal with ; a, of course, 
being Z a l r , b being 27 6 r , and I being 2 l r . There are, of 
course, 

aA 



ways of distributing the first type atoms in the manner 
indicated. 

&!_ __ 
bjbj.. .. . .& c ! 

ways of distributing the second-type atoms ; and 



ways of distributing the molecules. The product of these 
three expressions is then the number of different ways of 
producing the state indicated with the particular atoms and 
molecules chosen. But the italicised phrase shows us that 
we have by no means obtained all the complexions embraced 
within the state indicated. 

We have, in fact, to bear in mind that we can select a 
atoms from the given total number, a, of first-type atoms in 



or 



a ! (a a) ! a ! l\ 

02 



84 STATISTICAL MECHANICS FOR STUDENTS 
ways, and b atoms from the /? atoms of the second type in 

J*L 

bill 

ways. Hence the previous product must again be multi- 
plied by these two factors since a change of the individuality 
of a single atom in a cell yields a different complexion. Nor 
is this all. Having selected the dissociated atoms of each 
type, we have by no means settled the individuality of the 
molecules although we have, by reason of the selection 
mentioned, chosen the 2 I atoms out of which they are 
to be constituted. Let us consider a definite complexion 
obtained as indicated. Without altering the individualities 
or numbers of the atoms in the cells, and without altering 
the numbers of the molecules in the cells, we can obtain I ! 
complexions from this one complexion by permuting the 
atoms in the molecules. It is just as if we had dumb-bells 
each made with a red and a black ball arranged in a par- 
ticular scheme. Imagine the red balls to be removable ; 
there are I ! different ways of arranging them in the I places 
of the scheme and attaching them to the fixed black balls. 
Thus to obtain the total number of complexions consistent 
with the state described, we have to multiply the product 
of the previous five expressions by II. The final result is 



_ ^^ 

(8.1.3) 

The calculation of the most probable state proceeds as 
before by taking the logarithm of this expression and varying 
it under the assigned conditions of number and energy. 
Using the Stirling approximation, we obtain the following 
variational equations for the state of maximum probability 

Z log a r 8a r + 2 log b r 8b r + Z log l r 8l r = 
Z $a r + Z U r = 
Z 8b r + 2 81, = 
Z ar 8a, + Z br 86 r + Z lf $l r = 



CHEMICAL EQUILIBRIUM IN A GAS REACTION 85 

These, by the method of undetermined multipliers, give 
us 3 c equations 

log a r + A a + M *ar = 

log b r + A 6 + fi c* = . . . .(8.1.4) 
log l r + A a + A, + p, lr = 
These equations combined with 



are sufficient to determine the 3 c + 3 quantities, A fl , A 6 , p, 
and the a r , b r) l r in terms of E, n and the parameters. The 
usual feature of a common distribution constant appears 
once more and in the normal state the numbers are 
a r = A exp(-p. ar ) 



l r = A B 
where A = exp(-)( a ) 



The well-known mass-law follows at once from this. For 
the total number of atoms of the first type in the normal 
state is given by an integral (after the manner of Chapter 
IV.) such as 

= -J ...... Jea?p(-fu: a )i 

_ A /2m a 77\i 
~~ g \ ft / 



\dv 



_B /2jr 

flf \ ft 
== AB /2(m. + 

gr V /i 



Thus, since the concentrations are proportional to the 
n^mbers, we have the usual law 



86 STATISTICAL MECHANICS FOR STUDENTS 

where K is the equilibrium -constant depending on the 
temperature and the energy. Further, since K oc e~* w 



= - 

R0 2 ' 

where Q is the heat of dissociation of a gram -molecule of the 
molecular compound. 

8 . 2 Reactions in which the Number of Molecules is Un- 
changed. Turning to a reaction in which two molecules of 
the type XY yield one molecule of the type X 2 and one of 
the type Y 2 symbolised by 

2 XY ^ X 2 + Y 2 , 

we have in a particular state I molecules of XY, a of X 2 and 
b of Y 2 , so that if a is the total (and given) number of atoms 
X, and j3 the total number of atoms Y present, then 

2a " M==a (82 n 

26+Z=j3 * ' ' (8 ' 2 - 1} 

Each molecule is characterised by a definite amount of 
internal potential energy over and above its kinetic energy 
as a whole,* so that 



4 m a 

_e + ^ + 
& w>h 



where w a , w b , w t are definite amounts of energy, positive or 
negative. 

The counting of complexions proceeds on much the same 
lines as before. Considering a state as defined in section^ 

* We are excluding rotational energy. This could be treated by fresh 
co-ordinates and extending the dimensionality of the phase -diagram. It 
would complicate the mathematics without adding anything vital to the 
discussion at this point. 



CHEMICAL EQUILIBRIUM IN A GAS REACTION 87 

(8.1) we have, so long as we particularise the molecules, to 
obtain the number of complexions by multiplying 

a! 6! /! 



But for any complexion we can select the 2 a atoms in the 
X 2 -molecules in 

a! a! 

or 



2 a ! (a 2 a !) 2 a !Z! 

ways out of the a X-atoms ; and out of the /J Y-atoms we 
select the 2 6 Y 2 -molecules in 

_L 

2b\l\ 

ways. As before, this does not exhaust the possibilities. In 
any choice we are left with I particular X-atoms and I par- 
ticular Y-atoms to make the I XY-molecules ; but out of 
any complexion with particular XY-molecules we make I \ 
new complexions by keeping the X-atoms of these molecules 
fixed, as it were, in the design and permuting the Y-atoms 
in all possible ways ; for thus we form new molecules with 
the same atoms as before. We must be careful in applying 
a similar process to the X 2 -molecules. Out of any com- 
plexion with a particular X 2 -inolecules, it would look as if 
we could make 2a ! new complexions by permuting the 2a 
atoms all possible ways in the design ; but this gives us too 
large a result ; for many of these would only be replicas of 
other complexions (in the same state) made with the same 
a X 2 -molecules. Any permutation which did not destroy 
the companionship of the pairs of atoms would do this, and 
this fact shows us that the complete permutations suggested 
would reproduce definite complexions over and over again 
as many times as there are permutations of the a molecules 
(each regarded as a unity) in the scheme. But this number 
of times is a !, and so out of the particular complexion with 
a particular X 2 -molecules, we can only produce 2 a I/a ! new 
complexions by permutation of the atoms. A similar 
remark applies to the Y 2 -molecules. Thus the product of 
the five expressions quoted above must be multiplied by 



88 STATISTICAL MECHANICS FOR STUDENTS 

I !, 2 a I/a I and 26 !/6 ! to obtain the number of complexions 
possible for the state mentioned. It is 



ajaj ...... ailbjbjt ...... b c \ IJ. lj ...... l c \ 

as before. The normal state is obtained by the variational 
equations of the type 

Z log a\ 8a r + Z log b' r Sb r + 2 log l r 8l r = 
2 Z 8a r + 2 8l r = 
2 Z 8b r + 2 SZ; = 
2 ar Sa r + 2 e br Sb r + Z er Sl r = 0, 

leading to 3 c equations such as 

log a r + 2 A a + n * ar - 
log b r + 2 A, + p br = 
log l r + (A fl + A 6 ) + p, e lr - 

and three further equations depending on total numbers and 

energy. Thus in the normal state 



. . (8.2.3) 
l r = ABexp(-p, lr ) 
where 

A == exp(-\ a ) 
B = exp(-Xb) 

This result, combined with (8.2. 2), leads directly to the 
result that 

& 6 ____ T7" 



where K is the equilibrium constant and is proportional to 

t* ( w a 



a b 

expression w a + w b 2 w t is the energy absorbed in 
the reaction 

2 XY -> X 2 + Y 2 

and yielded in the reverse reaction, so as before 
eZlogK_ Q 



d8 
where Q is the heat of reaction at temperature 6. 



CHEMICAL EQUILIBRIUM IN A GAS REACTION 89 

8 . 3 Generalisation for Any Type o! Reaction. The method 
of procedure for any type of gas reaction is now fairly obvious. 

Let there be in all a atoms of type X, j8 atoms of type Y, 
y atoms of type Z, and so on. In the reaction there occur 
molecules containing x l of the X-atoms, y l of the Y-atoms, 
z I of the Z-atoms, etc., #i, y l9 z ly etc., being positive integers ; 
these we shall call Li-molecules. There will also be La-mole- 
cules, containing x 2 of the X-atoms, y 2 of the Y-atoms, 
z 2 of the Z-atoms, etc., and so on. The reaction follows 
some stoichiometric equation, such as 

v l LI + v 2 L 2 + v 3 L 3 + ...... =0 

where v x , i> 2 , v 3 , ...... are integers, some positive, some 

negative. This being so, the following equations must be 
true 



v l yi + V 2 2/2 + "3 2/3 + ...... (8 . 3 . 1) 



The energies of the molecules in a given phase of position 
and translational momentum are for an Li-molecule 



for an L-molecule 



and so on, where w l9 w 2 , ...... are definite internal energies, 

and m v m 2 , ...... are masses of the respective molecules. 

The heat of reaction during the reaction of v Li-molecules, 
v 2 L 2 -molecules, etc., is 

*i w i + "2 W 2 + V 3 ^3 + ....... (8.3.3) 

If now in any state there are present l^ Li-molecules, Z a 
L^-molecules, etc., then 

X l l l +x 2 l 2 + x 9 l 3 + ...... == a 

yi'i + yi'i + y a 'a + ...... =P (8.3.4) 

z l l l + z 2 l 2 +Z B 1 3 + ...... = y 



90 STATISTICAL MECHANICS FOR STUDENTS 

The number of complexions in which there are 
l n Lj-molecules, J 21 L 2 -moleeules, Z al L 3 -molecules, ...... 

in cell 1 
Z 12 Li-molecules, Z 22 L 2 -molecules, Z 32 L 3 -molecules, ...... 

in cell 2, etc., 

is, after a series of steps similar to those taken previously, 
found to be 



The variational equations are 

-Slog ^ SZi, + log Z 2r 8Z 2r + ...... - 

x l 2 81^ +x%Z 8l 2r + ...... = 

Vl Z Sl lr + y z 2 8/ 2f + ...... = 

+zZSl ...... -0 



2 lf 8^ + 2 c 2r 8Z 2f + ...... - 

leading to equations such as 



I 



2r 



and thus we find that in the normal state the numbers of 
each type of molecule in the r th cell are given by 

l lr = A*> B* (? exp(- Me lr ) 

(8.3.6) 



where 

A = exp( A ), B = exp( A 6 ), ...... 

As before, integration over the phase-diagrams gives the 
total number of each kind of molecule present in the normal 
state, and we find for the separate concentrations 



C 2 = M 2 A 35 - B^ (? ...... exp( - /W 2 ) . (8.3.7) 



CHEMICAL EQUILIBRIUM IN A GAS REACTION 91 

where Mj, M 2 , ...... are numerical multipliers arising in 

the integrations. Owing to equations (8 . 3 . 1) it follows that 
Cj" 1 C a " ---- = M exp { - j* (i w : + v z > a + . . .)} 



where M is equal to M/ 1 M. 2 V * . . . . 

and Q is the heat of reaction per gram molecule of re- 
actants or resultants. Thus the equilibrium constant 
makes its appearance again as a quantity proportional to 
exp ( Q/R0) leading to 

dHogK^Q^ 
dd R0 2 ' 



CHAPTER IX 

INTEBMOLECIJLAB FOBCES 

9 . 1 The Effect of the Finite Sizes of Molecules. In 

deriving the formula for the pressure of a gas in section (4 . 2) 
it was implicitly assumed that we were dealing with a swarm 
of point-particles. But the impossibility of crushing a body 
of liquid or solid into an infinitesimal volume is simple and 
direct evidence that whatever be the structure of molecules, 
they have finite size in the sense that the centres of two 
molecules cannot be forced nearer to one another than a 
certain definite distance, minute though it be. Contrasting 
two molecular systems, therefore, each one containing the 
same number of molecules in the same volume, but one being 
constituted of larger molecules than the other, it will be 
seen that the mean free path between collisions will be a 
trifle shorter in the first case than in the second. This will 
have the effect of slightly increasing the rate of transfer of 
molecules across an element of surface, thus producing at 
the same temperature, i.e., at the same mean velocity, a 
somewhat enhanced rate of transference of momentum. 
The effect will be all the greater the larger the molecular 
size in comparison with mean free path, i.e., the larger the 
concentration. We may, therefore, infer on general grounds 
that the pressure of a gas whose concentration is v molecules 
per unit volume is given more accurately than before by 
some formula, such as 

p = v iff (v) kO, 

where (v) is a function which approaches unity as v 
approaches zero, and increases in value as v increases. 
Expanding iff (v) as a series in ascending powers of v, we can 
as a first approximation retain the first power of v only and 
write 

2> = v(l +0v)&0 . . . (9.1.1) 

92 



INTERMOLECULAR FORCES 93 

as a somewhat amended form of the simple Boyle's law. 
From the reasoning employed, it will be realised that as 
between different gases, the constant /J will be larger for 
larger molecules. We have in the reasoning implicitly 
idealised a collision as an instantaneous phenomenon. It is 
scarcely probable that the actual occurrence is dynamically 
so simple ; still it is evident from the cohesion of solid and 
liquid matter and the broad facts of their compressibilities, 
that the repulsive forces called into play at the close en- 
counter of two molecules disappears at a very small distance 
apart, and so the conversion of kinetic energy into potential 
energy during the encounter occupies a very brief time as 
compared with the mean interval between collisions. It is 
this fact concerning the intermolecular repulsive forces 
exerted at very near approach which allows us to dispose of 
them in the somewhat cavalier manner employed above. 

9 . 2 Intermolecular Attraction. The elementary applica- 
tion of the conception of molecular attraction to the explana- 
tion of latent heat of vaporisation of a liquid is no doubt 
known to the reader. In the interior of the fluid a molecule 
does not experience a constant force in any definite direction, 
as it is surrounded by molecules whose resultant pull on it 
will be on the average zero. Only in the molecular layer at 
the surface whose thickness is equal to the radius of mole- 
cular attraction (beyond which the force becomes negligible) 
will there be a resultant pull normally inwards. This can be 
regarded as equivalent to removing the molecular attraction 
and replacing it by an increased external pressure. The 
amount of this increase is not difficult to estimate. Let/ (r) 
represent the magnitude of the force between two molecules 
separated by a distance r. Consider a particular molecule, 
A, and the molecules surrounding it in a spherical shell 
between spheres of radii r and r + 8r. Suppose the fluid to 
expand uniformly by a small amount so that there is a linear 
coefficient of extension of value e ; as the mutual potential 
energy of two molecules will increase by the amount / (r) er, 
it follows that the increase in the mutual potential energy 
of A and its neighbours in the shell will increase by 

rf(r) vlnr 2 Sr. 



94 STATISTICAL MECHANICS FOR STUDENTS 

Hence the increase in the mutual potential energy of A and 
all its neighbours is given by the integral 



4:776 V 



|>/(r)dr 



where or is the nearest distance of approach of molecular 
centres. The increase in the whole potential energy of 
attraction of the molecules will be obtained by summing this 
for all) the molecules and taking half the sum. (Otherwise 
the mutual energy of any pair of molecules would be counted 
twice.) The result is 

2 TT e n v 

where n is the total number of molecules, so that v = n/v. 
But if an increase of volume 8v accompanies the linear 
extension , then e = | Sv/v, and so the result for the 
increase in potential energy is 

a v 2 8v, 
where a = ~ ("r 3 f(r) dr . . . (9.2.1) 



If the work thus done in the expansion had been performed 
against an external pressure &>, instead of against molecular 
attraction, its value would have been to o> Sv. Thus the 
change of momentum produced in the molecules by the 
cohesion exerted in the molecular layer is the same as that 
produced by an external pressure of amount a i> 2 . This 
correction shows that the internal pressure, i.e., the rate of 
transference of normal momentum across unit area in the 
interior of the fluid, is p + a v* and not p. So, recalling the 
correction made in the previous section, we have 

p -fa v 2 = v(l +pv)k9 . . (9.2.2) 

If we write a = a tit and b j8 n, we obtain as a better 
approximation to the equation of state than Boyle's law the 
following result 

. a nlcO /. . b 

p +- = 



INTERMOLECULAR FORCES 95 

or putting (1 + bjv)~ l approximately equal to 1 bjv 

v-b)=KO (9-2.3) 



This is Van der Waal's famous equation, and of the various 
interesting conclusions to be drawn from it the reader can 
inform himself in text-books of physics or physical chemistry. 
We are here concerned with the effect produced on our 
statistical methods by the introduction of intermolecular 
force into the arguments, and we shall therefore concentrate 
on one result which follows from a study of the rela- 
tion (9.2. 3). 

It is well known that if isothermal curves are drawn, 
using Van der Waal's equation, these curves show a charac- 
teristic feature when the temperature is low enough. Travel- 
ling along such an isothermal in the sense of increasing 
volume, the pressure diminishes to a minimum, then in- 
creases for a space, reaches a maximum value, and once more 
proceeds to dimmish indefinitely. There are, in fact, two 
values of v where dp/dv is zero, and between them dp/dv is 
positive. Within this region of pressure there are, in fact, 
three values of v mathematically possible for each value of 
p (9, of course, being given).* The one which lies on the 
part where dp/dv is positive is considered to be so physically 
unstable as to escape observation by reason of its transience 
if it were produced. Of the other two, one is considered to 
correspond to a vapour phase which may be in an absolutely 
stable unsaturated or saturated state, or in a less stable 
condition of supersaturation ; the remaining value of v is 
associated with the liquid state, which in its turn may be in 
a relatively unstable superheated condition, or quite stable 
below its boiling point at the pressure. This interpretation, 
of course, implies that the temperature below which the 
particular form of the isothermals manifests itself is the 
critical temperature of the substance. In this way is the 

* Since (9 . 2 . 3) is for given p and a cubic equation in i>, there are, of 
course, three values of v possible for any value of p, whether within the 
range specified or no. But, of course, two of these may be imaginary, or 
if all are real, they will, if not in the range, lie on those parts of the curve 
(which, be it noted, has really two branches, one not being usually shown 
in the books) for which v < 6 with which we are not physically concerned. 



96 STATISTICAL MECHANICS FOR STUDENTS 

equation (9.2.3) linked up with the hypothesis that there 
is a continuity between the liquid and gaseous states of 
aggregation through intermediate homogeneous states which, 
however, although physically conceivable, could only have a 
very transient existence if actually produced . Before passing 
on to consider this fact from the point of view of the statis- 
tical methods employed in previous chapters, we may 
realise its possibility in a general way, apart from special 
analysis, as follows. The internal pressure, R0/(v 6), is 
reduced by the cohesion, a/v 2 , in the surface-layer to the 
value, p, of the pressure actually observed. An increase in 
volume implies a diminished density, and therefore a decrease 
in the internal pressure, i.e., a smaller rush of molecules 
across the inner surface of the molecular layer. But the 
diminished density also produces a diminished cohesion, 
and therefore less hindrance to these molecules in escaping 
across the outer surface of the layer, so that the net result 
might be actually a greater external pressure. Students, 
through a failure to grasp the real meaning of the symbol p, 
sometimes fall into the error of imagining that a homo- 
geneous state of aggregation in which p increases with v, 
is physically impossible, " because,' 5 as they say, " it is 
impossible for the pressure to increase if the volume in- 
creases," thereby betraying the fact that they think that 
the symbol^ refers to the internal pressure, concerning which 
the statement is true enough. The very great instability of 
such a state (not its impossibility in an absolute sense) can, 
however, be realised by conceiving it to exist and then 
considering what would happen if by reason of the mole- 
cular motion a small fluctuation began in the density of a 
small portion of it. In the ordinary way, if a small portion 
of a fluid expands, there is a reduced outrush of molecules 
from its original volume. The surrounding fluid is pouring 
in molecules at the normal rate, with the result that further 
fluctuation in this direction is checked, and a similar con- 
clusion follows for a fluctuation involving transient increase 
of density. But in the state of aggregation imagined, this 
would not take place ; the outrush is accompanied by a 
reduction in cohesion of such magnitude that the fluctuation 



INTERMOLECULAR FORCES 97 

is not checked, but actually assisted to greater intensity, as 
the increasing outrush overwhelms the normal stream of 
molecules inwards. Thus would be set up an expansion of 
the original fluctuating element to a less dense and more 
stable state. Similarly an original fluctuation in some ele- 
ment, beginning with an increase of density, instead of being 
checked by an enhanced outflow of molecules from it, would 
be assisted on account of a too great increase in cohesion, 
and a consequent inhibition of the balancing outward stream, 
so that the normal inward stream from without would force 
molecules into the element and so produce a stable con- 
densed phase therein. In some such way we can visualise 
the separation of the unstable phase into droplets of liquid 
and a saturated vapour phase. 

9 . 3 The Probability of a Macroscopic State when Inter- 
molecular Action is Involved. The change introduced into 
the analysis of Chapter III. by the assumption of inter-mole- 
cular force is produced by the fact that the energy of a state 
specified by the numbers n l9 7i 2 , ...... , n c of representative 

points in the phase -cells is no longer a linear function of those 
numbers, as was the case when we wrote E equal to 2 r n r , 
the r being functions of the parameters of the system. If 
we write a general function / (n l9 n%, ...... n c , a v a 2 , 

...... aj, or briefly/ (n, a) for the energy of the state, the 

variational equation (3.2. 10) in section (3 . 2) must be 
replaced by 



leading to the result that we must determine the values of 
the n r in the most probable state as well as the functions A, fj, 
in terms of n, E, and the parameters by means of the c + 2 
equations 



2 n r = n 
f (n, a) = E 

As regards the function, / (n, a), it will, apart from the 
linear terms in n r (which involve kinetic energy and potential 

S.M. H 



98 STATISTICAL MECHANICS FOR STUDENTS 

energy arising from external bodies), depend on the quantities 
Z (n r ) 8 , this summation referring to the sum of the numbers 
n r over those phase-cells which have one space-element in 
common. If the c phase-cells are constituted by associating 
each one of a elements of volume with each one of /? elements 
of extension-in-momentum (c being therefore equal to a /?), 
then this summation is over the j8 phase-cells which have 
the 5 th element of volume in common. Calling the sum N,, 
we have 

c / 

/ (n, a) = S r n r + </r (N^ N 2 , .... N a a l9 a 2 . . . . aj, 

r-l 

r being the sum of the kinetic and external potential 
energies in the r th phase-cell. Hence 

df^a) ^ (N, a) 3N, 

dn r r 3N, dn r 

a (N, a) 
~~ 9N. " 
= *, +Xs (Ni> N 2 , NJ. 

Here x (N 1? N 2 , N tt ) can be regarded as the increase 

in mutual potential energy produced by introducing one more 
molecule into the element of volume which is a constituent 
part of the r ih phase-cell. If we assume that the phase-cells 
are constructed in such manner that their constituent volume 
elements are considerably larger than the sphere of mole- 
cular interaction (which from what we know of the range of 
molecular forces is not at variance with the assumption of 
the physical smallness of those elements), we can take 

^, (Nj, N 2 , NJ to be the mutual potential energy of 

one molecule with respect to all the rest in the s th volume- 
element. It will, therefore, be a function of the concentration 
of the molecules in this element which we shall denote by v t ; 

so we write, instead of ^ (N x , N 2 , NJ, the functional 

form <f> (v g ) of the one variable v 8 . If we could assume some 
simple law of force for the attractions, and if we could assume 
this law to hold down to any distance apart (virtually 
assuming the molecules to be point-centres of force), <f> (vj 
would have the form a v^ a being a positive constant, and 



INTERMOLECULAR FORCES 99 

the minus sign being due to the fact that for attractive 
forces potential energy decreases with decreasing separation 
of molecules and increasing concentration. But, of course, 
this overlooks the occurrence of intermolecular repulsions 
at close encounter. The fact that there is a finite limit to 
the value of v under the greatest pressures conceivable, 
requires us to assume that although <f> (v) may very well 
behave as a v for relatively small values of v, yet for 
values approaching some maximum limit v , <{> (v) must 
begin to increase in value not only up to zero once more, but 
actually to positive infinity if we are to regard concentrations 
beyond v as physically impossible. This is in fact the 
counterpart of the infinite value required for the pressure p 
in Van der Waal's equation to reduce v to the value h. 

From these equations we now find for the number of 
molecules with representative points in the r th phase-cell, 
in the most probable state the value 



This, by summation over the /? elements of extension-in- 
momentum, which linked up with the s ih element of volume 
yield /? phase-cells, gives N, as 

ft 

C i*.<b(v K ) y 0ntf /Q o i \ 

6 ' Z/ 6 ~ , . , . ^y . O . 1 j 

N 8 , if divided by the magnitude of the element of volume, 
gives v g , and since the summation can be as usual replaced 
by an integration over all values of momentum, we obtain 
finally the result 

+ 00 -f QO + 00 

Vg == D e-^W J J J e ""^ d dr ) d $ 

** (V J . . . (9.3.2) 

where D is a constant. We are disregarding any external 
field of force. 

From (9 . 3 . 1) it follows just as before that 3/(2 p) is the 
average kinetic energy of a molecule, in any element of 
volume, for the factor e"" M * ( " ) , being the same for all phase- 

H 2 



100 STATISTICAL MECHANICS FOR STUDENTS 

cells associated with that element of volume, does not invali- 
date any of the steps occurring in the calculation. Thus 
internal potential energy does not interfere any more than 
external with the validity of the law of energy-partition. 

The result (9.3.2) shows that the most probable state is 
consistent with a uniform spatial density given by any root 
of the equation in 

v = A /J e~^ (v) , . . . (9.3.3) 

where A is a constant. 

The easiest way to see how many roots are involved is to 
plot two graphs 



y = e -^ ( *> (9.3.5) 

the co-ordinate x representing the concentration v. The roots 
of (9 . 3 . 3) are the abscissae of the points of intersection of the 
graphs (9.3.4) and (9.3. 5). Now if, as suggested by the 
neglect of forces of repulsion, <f> (x) were taken to be a x, 
there would either be two roots or none, according as the 
straight line y = &/A/J, cut or did not cut the exponential 
curve, y = e* ax . But if we take account of the finite size 
of the molecules, <j> (x), although increasing in proportion 
to x at first, will, as x gets very near to v in value, pass 
through a maximum and decrease through zero to minus 
infinity at x = v . Hence the curve, y e~^ (x \ will at a 
certain point in its exponential ascent abruptly turn down 
and y will decrease to zero at x = v . 




The curve y = e "" 



INTERMOLECULAR FORCES 101 

We thus see the possibility of the straight line (9.3.4) 
cutting the curve (9 . 3 . 5) in three points. If, however, //, is 
reduced in value, i.e., the temperature raised, the slope of 
the line will be increased, and the peak of the curve will not 
be so high, and so the line would only cut the curve in one 
point corresponding to a relatively small value of v. If on 
the other hand /x is increased, the line will fall low, the curve 
will reach high ; once more there will be only one point of 
intersection, this time, however, yielding a concentration 
close to v . We thus reproduce once more the result usually 
derived from Van der Waal's equation. The author has not 
met elsewhere this method of presenting the matter, but 
thinks it is perhaps worth while to outline it here, as it links 
up the result with the usual statistical method of estimating 
probabilities by counting complexions consistent with the 
energy conditions. However, in order to bring out the great 
instability of the state indicated by the point Q, we must 
appeal to a very ingenious piece of analysis first given by 
Smoluchowski in 1904. It arises in connection with the 
mathematical discussion of fluctuations in density of a 
molecular system. This will be dealt with in the next 
chapter. 



CHAPTER X 

FLUCTUATIONS OF DENSITY IN A MOLECULAR SYSTEM 

10.1 The Probability o! the Occurrence of a Prescribed 
Number of Molecules in an Element of Volume. Reverting 
to the fundamental formula of the theory, we find that when 
uniform concentration of molecules in a definite volume is 
the most probable state (there being no external field of 
force), the number of complexions associated with a state in 
which n l9 n 2 , ...... , n c molecules occur in the c elements 

of volume respectively, is 

_ n\ _ 

njnj. ...".. .n e \ . . . . (10 . 1 . 1) 

Hence the probability that there shall be a definite num- 
ber, Z, molecules in the r th volume element is proportional to 
the sum of expressions such as (10 . 1 . 1) for all the states in 
which n r I, i.e., to 

1 27 _ U - _ (10.1.2) 
l\ fti f . n 2 \ ...... n,^ n r ^\ ..... n c \ 

the summation being for all positive integral values of 
H!, n 2 , ...... ra r _i, w rfl , ...... n c consistent with 



If (n I) \ were substituted for n ! in the numerator, the 
terms in this summation would be the coefficients of the 
x a y ft z* ...... products in the expansion of 

(x + y + z + ...... toe 1 terms)""' 

Hence the expression (10 . 1 . 2) is equal to 



(n - I) \ I ! v 

But c is a large number, and the last factor is practically 
c w /c*. Also, as we know that large deviations of I from the 

102 



FLUCTUATIONS OF DENSITY 103 

average value n/c are rare, I is very much smaller than n 

and we can write n 1 for n (n 1) (n 2) (n I + 1). 

Thus (since c n is a constant), we find that the probability 
sought is proportional to 



n 1 a 1 



where a is the average value of the molecules in an element 
of volume. Finally, since the sum of these probabilities for 
all values of I is to be unity, and since the sum of the expres- 
sions a l /l ! for all values of I is e a , we see that the probability 
that there shall be / molecules in an element of volume, the 
average number being a, turns out to be 

e~* ..... (10.1.3) 

A simple scrutiny of the result shows that a l /l ! increases 
as / increases up to a, and thereafter decreases as I travels 
beyond a in value ; thus there is a maximum probability 
for / = a, as there should be, of course. The average value 
of the deviation I a is, of course, zero, but the average 
value of its numerical magnitude is not. It is, as usual, 
estimated by means of the average value of the square of 
I a. This will be 



_ a | la 1 

' s 



By a rearrangement of terms, this can be written 
,- r( 




104 STATISTICAL MECHANICS FOR STUDENTS 

Thus the mean deviation from the average value a of 
the number in the element is a*. Clearly this mean 
deviation becomes proportionally less as the concentration 
in the element increases. For a million molecules in an 
element on the average, the mean deviation is a thousand 
either way, or one in a thousand. For a hundred million 
molecules in the element, the deviation is one in ten thousand. 
This proportional deviation measured thus, is called the 
" condensation/' and it is clear that the average condensation 
in an element holding a molecules on the average is 1 in a* 
or a~*. 

Returning to (10 . 1 . 3), we see that the ratio of this expres- 
sion to its maximum value is 



a 1 la a 
/ 

/!/! 



a l -*a\ 
or 



l\ 
Taking I as greater than a, this can be written 



a. a. a. 



(a + 1) (a + 2) ...... I' 

there being j factors in numerator or denominator where j 
is the deviation I a. Hence the logarithm of this ratio is 



which = -- ...... - - + negligible terms 



2a 

where, as usual, we disregard as of no importance the states 
where j would become more than a small fraction of a. If 
I < a and j = a I, the ratio is 

a (a -r 1) I 

a. . a a 

and the same result emerges for the logarithm of this ratio. 

If we now wish to convert this result into a form suitable 

for dealing with continuous changes of density within a 

volume element rather than discrete changes in numerical 



FLUCTUATIONS OF DENSITY 105 

concentration, we replace j (j + l)/2a by ^ a y 2 , where y is 
the condensation in density, j/a ; for in this case unity will 
be negligible compared with the values of j since the average 
value of j is a* and is actually an enormous number even 
in a physically small element. Thus if the density p is 
given by 

P=fc,(l+y), 

where p is the average density, the probability that the 
density in the element lies between p and p -f- Sp, i.e., the 
condensation y between y and y -f- Sy is P(y) Sy, P(y) being 
a function which satisfies 



where P is the maximum value of P(y), occurring when 
y = 0. P is readily obtained from the fact that 

P(y) dy = I* 

-CO 

and by a reference to the Appendix to Chapter I., turns out 
to be (a/2 TT)*. Thus the probability that the condensation 
is in the range y to y + Sy is 



/ay 
\2W 



e * Sy . . . . (10.1.4) 



a being the average number of molecules in the element. 
The chance of a particular condensation is smaller the greater 
the value of a ; this we have already deduced, but it is once 
more evident from (10 . 1 . 4), the decreasing exponential 
factor quite easily swamping the increasing factor a*. 
From (10 . 1 . 4) we can once more calculate the average value 
of the condensation, i.e., the root-mean-square condensation. 
It is 



a \* f+ 

J k 



* The reader may think it absurd to integrate from negative infinity 
for 7, since by definition the numerically greatest value of 7, i.e. (I a)/a, 
on its negative side is unity when I ia zero. Fortunately we are rescued 
from this apparent absurdity by the fact that by the time ay 2 / 2 has reached 
a numerical value such as 10, the outlying parts of the integral are negli- 
gible, and on account of the great value of a this is attained by quite small 
values of 7 positive or negative. 



106 STATISTICAL MECHANICS FOR STUDENTS 

, . , / a \* 1 TT* 

which = I ) . - | 

\2W 2 (a/2)* 

_ 1 

> 

a 

so that the root-mean-square of y is a~* as determined earlier. 
Calling the mean square value y 2 , we can write (10 . 1 . 4) in 
the form 

^ exp { %= )Sy. 

27r/ V 2yV 

10 . 2 Smoluchowski's Theory of the Unstable States of a 
Fluid. The usual presentations of Smoluchowski's theory 
are so brief in the initial statements leading the funda- 
mental equation of the theory that there is an element of 
obscurity and vagueness in the mind of the beginner con- 
cerning its validity. It is hoped that the following preamble 
will remove such doubt. 

The fluctuations of density which occur throughout the 
body of a fluid owing to the molecular motion, imply that 
the volume of a given number of molecules is always varying. 
Instead, therefore, of considering the fluid as a molecular 
system, let us for the moment think of it as a continuous 
medium at rest in the broad sense, but pulsating throughout 
with compressions and rarefactions. Let us further conceive 
it to be divided into elements of volume, containing the same 
mass, which will therefore be equal if the density be truly 
uniform throughout. Let there be N of these elements, and 
for the moment let us consider that each of them is capable 
of having by reason of the pulsations in size any one of c 

discrete values of volume, v lt v 2 , v c . (We are here, 

just as in Chapters I. to III., compelled to adopt in the 
initial stages of the reasoning the standpoint of discon- 
tinuity which will be later modified to suit the requirements 
of continuity.) The reader will now easily realise that a 
complexion of this system will be specified by saying that 
Nj particular elements have a volume v l9 N 2 particular 
elements a volume v 2 , etc. The number of complexions is 
as usual N !/Nj ! N 2 ! N c ! for the statistical state 



FLUCTUATIONS OF DENSITY 107 

Nj, N 2 , , N c . As usual, this will be assumed to be pro- 
portional to the probability of the occurrence of the state. 
To find the most probable state, we proceed as before, 
taking account of the energy condition. This condition is 
obtained in the following manner. If v is the volume of any 
element at uniform density throughout, and p the uniform 
pressure then existing, the potential energy in an element 
of the fluid at volume v r over and above the energy at 
volume v n is 



(10.2.1) 



p being the pressure at volume v. To justify this, remember 
that if v is smaller than v 09 p is greater than p oy and so work 
would have to be done on the element to force it into a 
smaller volume than that consistent with a pressure p from 
its environment. (The case is analogous to that of a body 
suspended at the end of a spring. If the tension T at a 
length I is greater than the weight w, then work has to be 
done in forcing the body down to this length of the spring.) 
This work is the product of p p , and the decrease in 
volume, i.e., the product of p p and $v. If, on the other 
hand, v is larger than v and p less than p ot work still has to 
be done on the element to force it out into a larger volume 
than that consistent with the pressure p . (Again, there is 
the analogy of the body being higher than its mean position, 
there being a reduced energy of extension of the spring, but 
a more than compensating gain in the gravitational energy 
of the body.) The work done is the product of p p, and 
the increase in volume, i.e., (p p) Sv. As in the case of a 
body oscillating up and down there is increase of potential 
energy whether the body is above or below its mean position, 
so in the pulsating element of volume there is increased 
potential energy whether the size of the element is greater 
or less than v . The reader will naturally think of the 
accompanying changes of kinetic energy which would accom- 
pany such oscillations and pulsations in isolated systems ; 
but he is asked to bear in mind that we are counting com- 
plexions consistent with constant temperature throughout, i.e., 



108 STATISTICAL MECHANICS FOR STUDENTS 

changes in kinetic energy are ruled out by that proviso, and 
the energy condition to be satisfied is that gains in the 
potential energy of certain elements are to be compensated 
by losses in the others, or, in short, if we represent (10 . 2 . 1) 
by r 

N! X + N 2 e 2 + ...... + N c c = constant. 

Combined with the condition of a constant sum for the 
N f , we have the same solution as before. In the most 
probable state of the pulsating medium the number of 
elements whose volume will be v r at any instant is pro- 
portional to exp ( r] e r ), where 77 is a constant to be 
determined from the condition of constant number and 
potential energy. 

The equation of state of the fluid enables us to express 
(10 . 2 . 1) as a function of v r . (It may be as well to point out 
at this juncture, in view of the warning issued some pages 
back, that in (10 . 2 . 1) p is the external pressure ; the calcu- 
lation of work done on and by the element in its changes 
ensures that.) If a, 6, c, etc., stand for the partial differential 
coefficients dp/d v, d*p/dv 2 , d 3 p/dv 3 , etc., all estimated at the 
value v for t;, and, of course, at constant temperature, then 
we know by Taylor's theorem that 

P =Po +~j (V - V ) + - (V - V )* + - (0 -. t?,) s + ...... 



and thus by (10.2. 1) 






We can now adapt the preceding considerations to the 
idea that the volume of an element changes continuously, 
and not by discrete amounts. The number of elements which 
in the most probable state have a volume between v and 
v + 8v is given by 

D exp fa iff 
where 



o 1 



FLUCTUATIONS OF DENSITY 109 

and D and rj are constants to be determined from 

o 

* dv = N 



D ! 

Jo 

D 

* a 



e n *dv = constant. 



Instead of the variable v, we can introduce the condensa- 
tion, y, defined as (v - v)/v . The number of elements 
which in the most probable state have a value of condensa- 
tion between y and y + Sy is proportional to 



where 

</> (y) = Ay 2 + By 3 + Cy 4 + ....... (10 . 2 . 2) 

A, B, C, ...... being coefficients given by 



A- 

2! 



< 10 - 2 - 3 > 



If now we suppose that our attention is concentrated on 
one element of the fluid, we can assume the probability 
that at a given moment the condensation in this element has 
a value between y' and y" to be proportional to the number 
of elements which at any time have in the most probable 
state condensations within these limits. This probability is 
therefore proportional to 



J e&p 



We can find the value of 77 directly by a reference to the 
result (10 . 1 . 4) in the previous section. In the case of gases 
in a rare state for which one would anticipate that the 
analysis of that section holds, the molecules exerting no 
action on one another on the average, it appears that 77 <f> (y) 



110 STATISTICAL MECHANICS FOR STUDENTS 

should be equal to ay 2 /2, where a is the average number of 
molecules in the element of volume. Also as only relatively 
small values of y are effective in actual fact, we can ignore y 3 
and higher powers of y in the series for <(y) in the case 
contemplated. Hence 

77 A y 2 = - a y 2 /2 



02 

or T? - -- / < 

2 ! \3v 



_ a I ( ajc9 \ 
V/ \ T?"/ 



ak9 

P = 



J_^ 





Once more the distribution constant emerges and Smolu- 
chowski's result is written 



carp/ft <ji(y)Uy . . .. (10.2.4) 

The author trusts that the rather lengthy exposition is 
justified. Most accounts start baldly with the assumption 
that the probability in question is given by an expression 
involving 



-JL 
ke 



dv, 

r 

where E = I (p p ) dv, with no very clear reason given 

Jv 

for this value.* 

In order to apply the result to other states of the fluid 
than that of rare gas, we first of all note that if (dp/dv) is 
negative, the index in the exponential factor of the prob- 
ability is essentially negative for any value of y positive 
or negative, giving decreasing probability for increasing 
fluctuations, obviously a necessity of any stable state. If, 

* It will be perhaps wise" to warn the reader that E is more usually 
written 



f' 

a (p 

Jv 



po)dv, 



but that is because v is then the specific volume of the fluid, and so the 
factor, ma, the mass of the fluid in an element of volume is required. In 
the text v is the volume of the element. 



FLUCTUATIONS OF DENSITY 111 

however, (dp/dv) is positive in the state of uniform distri- 
bution and we have seen the theoretical possibility of this 
the exponent is positive for any value of y, giving increasing 
probability for increasing condensation or rarefaction ; in 
short, the state is highly unstable, and so the general in- 
ference drawn before is amply confirmed by this closer 
analysis. It will be remembered that on a Van der Waal's 
isotherm there may be two points where dpjdv is zero. If 
the average density of a fluid should correspond to either of 
these points, the series for c(y) would start with y 3 , and 
ignoring the remaining terms of the series, the probability 
would involve 



Thus, according to the sign of d 2 p/dv 2 , we would have 
increasing probability for fluctuations involving expansion, 
and decreasing probability for fluctuations involving con- 
traction or vice versd. This one-sided stability would be 
useless for maintaining a physically homogeneous state. 
Indeed, there are parts of the isothermals where dp/dv is 
negative, but which are so near to the points just mentioned 
that they can hardly be said to correspond to stability in a 
real physical sense, since they are associated with super- 
saturated vapour or superheated liquid. Smoluchowski's 
result is quite consistent with this ; for a negative but small 
value of A in e^ AY ' means that y must increase numerically 
more than usual before one goes beyond ordinary prob- 
abilities, and by that time the fluid in an element may, if 
the fluctuation has been in the suitable sense, have attained 
the highly unstable condition. 

But perhaps the most interesting application of Smolu- 
chowski's theory concerns the critical state at which dp/dv 
and d 2 p/dv 2 are both zero. In that case both A and B are 
zero, and the series for <f>(y) begins with Cy 4 . To determine 
the stability of this state we must determine a value for C. 
This can at all events be obtained approximately from Van 
der Waal's equation, which we write 

___ akO a 
P ~~ v 3 ~~v*' 



112 STATISTICAL MECHANICS FOR STUDENTS 

(We are using a for the cohesion constant, since a is being 
used as the average number of molecules in an element of 
volume, and for the minimum volume.) 
Thus 

24a 



W ~~ ( V - 0)4 
and by (10. 2. 3) 



v 4 (v 0) 4 

At the critical point it is known (since dp/dv and d 2 p/dv 2 
are zero there) that v = 30, p = a/270 2 and ak0 = 8a/270. 
Thus at that point 

a Slakd 



Slake 






8 64 

9a 



The series for <(y) now starts with y 4 , and neglecting higher 
powers, the probability now involves the exponential factor 

exp 

As this decreases with any variation of y from zero, the 
critical state is stable ; but closer investigation shows that 
the condensation fluctuates between wider values than in 
other conditions. This is proved by working out an average 
value for y. Thus the average value of the positive con- 
densations is (writing c for 9a/64) 



which, by writing x for c*y, becomes 

JOO 
e~ x * x dx 


{00 
e~* 4 dx 




FLUCTUATIONS OF DENSITY 113 

1 f 

The integral in the numerator is I eT^ dy, which is, 

2 Jo 

as we know, 7r*/4 ; the integral in the denominator can be 
calculated by quadrature to be 0-666 . . ., and so we obtain 
finally for the average value of the positive condensation 

1.13 



It is the appearance of the fourth root of a, and not the 
square root, as in the case of gases in a more usual con- 
dition, which is the interesting feature. Thus, in an element 
of volume containing 10 8 molecules, which for a substance 
in the critical state would have linear dimensions of the 
order of magnitude of the wavelengths of light, the value 
of this expression would be about *01, and so fluctuations in 
size of the order of 1 per cent, in density would be the 
average sort of occurrence. It is this result which is held to 
account for the well-known opalcscence which appears in a 
substance in the critical condition ; the real lack of homo- 
geneity in the medium is sufficiently marked to scatter the 
light of a beam which is passed through it. When the 
illumination is strong, a scattering of bluish light can be 
observed in a direction at right angles to the beam. Keesom 
has, in fact, linked up Smoluchowski's theory with physical 
optics, and deduced the well-known formula of Rayleigh 
for the intensity of the blue light of the sky, thus con- 
necting this phenomenon with fluctuation theory. 



CHAPTER XI 

THE SECOND LAW OF THERMODYNAMICS . II 

11.1 The Thermodynamical Equilibrium of a Condensed 
System. In Chapter VI. we deduced the second law of 
thermodynamics for a gaseous phase in statistical equi- 
librium. It is necessary to make sure that the deduction is 
still valid for systems in which we cannot ignore inter- 
molecular action. 

As we have seen in Chapter IX., the normal state of a 
system, for which the energy is given by a general function 

/ (n l9 n 2 , n c , a lt a 2 , a c ) of the numbers in the 

phase-cells and the parameters, is obtained from the equa- 
tions 

/ (n l9 n 2 , n c ,a l9 a 2 , a e ) = E 

% + ^ 2 + +n c = n . . (11.1.1) 

3f(n 9 a) 



and the c equations, log n r = A //,- 



dn r 



These may be regarded as c -f- 2 equations to determine 
A, E and each of the n r as functions of ft and the e parameters 



a l9 a> 2 > ...... 9 a e- We shall denote the functions so deter- 

mined by A (/A, a), H(fi, a), i/^fi, a), ...... , v c (p,, a). H(ft,a) 

is, of course, obtained from/(n, a) by inserting the c functions 
v r (ft, a) in place of the variables, n r ; i.e., 

H(ft,a) =/K,v 2 , ...... " c ,ai, a 2 > ...... > a e) (H.1.2) 

Let us also denote by x r (/^> a ) ^he function obtained when 
the functions ^ 1 (ft, a), *> 2 (ft, a), etc., are substituted for n l9 n 2 , 
etc., in df(n, a)/dn r ~ and by a (ft, a), the function obtained 
when the same substitution is made in df(n y a)/da s . 
It follows that 

dv, 



114 



SECOND LAW OF THERMODYNAMICS 115 

, 9H (u,(i) dv r , y / x /i-i i ^\ 

and ^ ' = Z x r i~- + O, ) - (H . 1 . 4) 

ca 8 r ~i ca s 

Recalling the general line of argument in Chapter VI., we 
consider a change from a normal state with values /*, a for 
the S.M. variables to a normal state with values fji + S//,, 
a + Sa. The change of energy 8E which is given by 

8a, . (11.1.5) 



ca 



is, as before, separated into two parts, one of which, SEj, 
is the change accompanying a variation in the parameters, 
but with the distribution in the phase-cells still left at the 
original normal distribution. The remainder is 8E 2 . It 
should be noted that SEj, the analogue of the mechanical 
work alone done on the system, is not given by the second 
term on the right-hand side of (11.1.5), as a glance at 
(11.1.4) will show. Actually 

, a) Sa, 

^_' S Zx r ^Sa. . (11.1.6) 



r . 
=i ca 8 r-i-i ca 8 

To proceed successfully from this point, our aim, as in 
Chapter VI., must be to discover a function ^(/^ a), such that 



8El = Z l Sa, . . . (11.1.7) 

=i da g 

and with that end in view, we write (11 . 1 . 6) as 

8E X = * j H^ a) - 2 X, v r \ Sa, + Z Z v, ^ Sa, 
8~ica 8 ( r =i ) r=i*-i va B 

(11.1.8) 

and endeavour to adapt the last term in this to the necessary 
form. Guided by the procedure in Chapter VI., we obtain 
from (11.1. 1) the following c equations, which are identically 
true for any values of the a e 

log v r (iJL, a) = A(^, a) p x r (fji y a) . (11.1.9) 

12 



,116 STATISTICAL MECHANICS FOR STUDENTS 

Differentiating this with respect to any of the a s yields an 
identity 

9*> 9A 9v, 

_ : = v r -- jLt V r 

da, da, da t 

and an addition of these gives 

9x, 9A dn 

a Z v r -^ = n since - = 0. 
r ~i ca s da, ca 8 

On substituting in (1 1 . 1 . 8), we find that (1 1 . 1 . 7) is valid 
if we define M* 1 (/x, a) thus 



This is obviously a generalisation of the definition in 
Chapter VI. ; for if f(n, a) is linear in the n r , f(n, a) = 
H n r df/dn r) and thus H (/x, a) = 27 x r ^ r ; so that in such 
case "^ reduces to n A//*. 

However, we have not carried the attempt to a successful 
conclusion yet. We must also see if the mathematical 
relation (6 . 2 . 7) in Chapter VI. is still valid. Naturally we 
follow the same procedure as we did there and differentiate 
the identities (11.1.9) with respect to p, thus obtaining 



^v---vx- v 
Sfji dfji 

Addition of these c relations yields 

nf-=Z v r Xr +M Zv* 

Cfl r ~l r-1 C7jLt 

From (11.1.10) we obtain 



, a) A 9 Xr rv 1v ndX ^A 

^/ v r ^j \ r -f- - - n ~. 



- - ----- - - r r - 

Cp, CfJL r=l OfJl. r=-l Of* (lOfJ. H 

which by (11.1. 3) and (11 . 1 . 11) 



" 1 A A 

= - 2 v, X f - - a , 

/A r=l i 



and by (11.1.10) 



^ _ .(11.1.12) 



SECOND LAW OF THERMODYNAMICS 117 

Thus, as before, 



and 

M SE 2 = 8[/i{(H/i,a) -(/*, a)} ]. 

The general line of argument proceeds as before to a 
deduction of the second law with the entropy defined by 

4>(/x,o)=fc/i{H(^o) -(,*, a)}, 
for the relations (11.1.9) give us 

c 

log W w = ^ log n E v f log v r 

r=l 

c 

= 7i log n n X -\- fji S v r x r , 

r=l 

so that 

* (log W m - n log tt) = fyijHGu, a)- ^(/x, a)} 



and the connection between increasing entropy and increas- 
ing probability is once more established. 



CHAPTER XII 

THE STATISTICAL-MECHANICAL THEORY OF A LIQUID AND A 
VAPOUR PHASE IN CONTACT 

12.1 Deduction of Clapeyron's Equation. An enclosure of 
given volume is supposed to have a portion of its volume V l 
occupied by a saturated vapour and the remainder V 2 
occupied by the liquid ; the number of molecules in the 
vapour is n l and n 2 in the liquid, so that the concentrations 
are given by v x = n l /V l and i> 2 = n 2 /V 2 . In Chapter IX. 
it was demonstrated that under these circumstances 

-^i) . . . (12.1.1) 

v ' 



\27rm 

v 2 = T> (-V 

\277-ra 

where </>(v) is the potential energy of one molecule due to 
intermolecular force in a place where the concentration is v. 
This result is not dependent on any particular functional 
form for </> ; but it will be found as we proceed that we must 
have a modicum of information about it in order to come to 
a definite conclusion in the problem on which we are engaged. 
First of all we know that <j)(v) approaches a maximum value 
as v decreases, and beyond a certain concentration it is 
practically constant ; so that in the gaseous phase we shall 
assume that <fr'(v) is zero, where <j>'(v) is written for d<f>(v)/dv. 
As v increases and the molecules approach one another on 
the average, the potential diminishes so long as the forces 
are attractive. But when the liquid state is reached, the 
average separation is such that there is a compensation 
between attractions and repulsions on the average, and 
further compression would involve an increase of the 
potential energy due to preponderance of repulsive force. 
So that the concentrations in the liquid state would be such 

118 



LIQUID AND VAPOUR 119 

that for any value of v in the narrow range involved, (f>(v) 
would practically be a minimum, so that <f>'(v) would be 
practically zero in this case also. Thus in order to give the 
necessary definiteness to our problem, we shall assume that 
the function < satisfies for the two phases in question 

f(*i) = f(" 2 )=0 . . . (12.1.2) 

By reason of (12 .1.1.) 

Vl e^h) = ^ e ^(v,) 
or 
log ft! +'/^K) log V a = log n 2 + jot^(v a ) log V 2 . 

Now consider the system with a slightly altered distri- 
bution constant, i.e., temperature ; this will involve an 
alteration in the numbers and concentrations as well as in 
the volumes V t and V 2 . Differentiating with respect to the 
temperature 6, and remembering that dfijdO = 1/&0 2 , we 
obtain 



__ . > 

! dd V l dd "*" w {Vl) dd 



^ _ _ _ 

i ~~n 2 d9 V 2 dd ' 2 d9 

so that, on account of the hypothesis we make as regards an 
ideal liquid and vapour state in (12. 1. 2), 



^. 

n! dd n 2 dO V l dd V 2 dO 

(12.1.3) 

We can simplify this result very markedly by assuming 
that the two phases occupy equal parts of the enclosure at 
the temperature 9. Since the enclosure has a fixed volume 
dVJd9 = dV 2 /d9, and if we now also assume that Vj is 
equal to V 2 , the two terms involving these quantities dis- 
appear. Further under such circumstances 

ni^v* 
n 2 v l 

where v l and v 2 are the specific volumes of the two phases, 
since n v l m and n 2 v 2 m are equal respectively to V l and V 2 . 



120 STATISTICAL MECHANICS FOR STUDENTS 

Lastly, since the total number of molecules is constant, 
dnjdd = - dn 2 jd0 9 and so the left-hand side of (12 . 1 . 3) 
becomes 



i l nj dO ' 
which 



( 1 v 2 \ I dn l 
"TjH^W 



v z \ d log K! 



On the right-hand side of (12. 1. 3), <t>(v^) <f>(v 2 ) is the 
energy required to remove one molecule from the liquid 
phase to the vapour, and since the volume of the enclosure 
is fixed, and therefore no external work performed, this is 
the internal heat of vaporisation per molecule. Denoting it 
by w, we have, as the final result 

... (12.1.4) 



and this leads directly to Clapeyron's equation, for if p is the 
vapour pressure, 

p = j/i k6 

and dlogp _d log Vl 1 

~~dO dT~ + 

v * w i ^ 

"" 



where L t - is the internal latent heat of n molecules and R = 
nk. Since the first phase is considered to be an ideal gas, 
R0 == pv ly and so 

dlogp __ Vi LJ + p(v l v 2 ) 



_ l 
== 



LIQUID AND VAPOUR 121 

where L is the ordinary latent heat. As v 2 is small compared 
to v v we have approximately 

d log p __ L 
~ ~ 



but keeping the exact equation, and once more writing 
for R#, we obtain 

d log p _ L 
d9 p0(v L v 2 ) 

d ?=- _ . . . (12.1.5) 
de 6(v l - v z ) v ' 

which is the proper result, and as a matter of fact, is 
the exact equation deducible by strictly thermodynamical 
reasoning for real liquids and vapours. 

12.2 The Relation between the Latent Heat and the Specific 
Heats. The second thermodynamic equation for a liquid- 
vapour system can now be easily obtained. Let s l and s 2 
be the specific heats of the vapour and the liquid, it being 
understood that these are the thermal capacities of unit 
mass of the vapour or liquid as heated in the fixed enclosure 
(not subjected to a constant pressure nor on the other hand 
with each portion maintained at constant volume, the usual 
conditions). The proviso is well known to those acquainted 
with the thermodynamical treatment of this matter. The 
vapour, for instance, is maintained in a saturated condition 
during the heating, involving a diminution of volume, and 
on that account it may happen that if the external work 
thus performed on the vapour is too great, heat would have 
to be removed from the vapour, and s l would be negative. 
This is in fact the case with water. In the case of the liquid, 
s 2 differs but little from the usual specific heat under constant 
external pressure. 

If now unit mass of the liquid is heated through 86, the 
heat supplied is s 2 86 y and external work p8v 2 is performed. 
Thus the internal energy of the unit mass of liquid is greater 
by s 2 86 p8v 2 . If unit mass of the vapour is similarly 
heated, its energy increases by s^O - pdv v (As just 
mentioned, 8v l is in general a negative quantity if 86 is 



122 STATISTICAL MECHANICS FOR STUDENTS 

positive.) Hence the energy difference between the liquid 
and vapour phase increases during the change of temperature 

by 



But this energy difference is L^ at first, and after the tem- 
perature rise L; -f 8L, t , so 



dd d0 

Hence 

d{L, + P(VI v 2 )\ . 

de - - = s *~ s * + 

and thus, by Clapeyron's equation 



To be sure by our hypothesis (12.1.2) 
dw 



__ A 

and L f will not change with temperature at all for our ideal 
liquid, which is a rather restrictive hypothesis. 



CHAPTER XIII 

THE SOLID STATE CONSIDERED AS A SIMPLE LATTICE OB 1 
MASSIVE PARTICLES 

13 . 1 The Specific Heat of a Monatomic Solid. We now 

come to the treatment on classical lines of the last topic to 
be dealt with before an endeavour is made to introduce the 
reader to the modifications of statistical-mechanical theory 
occasioned by the quantum hypothesis. On classical lines 
it lends itself to very simple treatment indeed. 

The rigidity of a solid, which is its characteristic feature, 
we shall idealise by conceiving it as constituted of a group 
of n particles, each of mass m, situated at the points of a 
simple space-lattice, so that choosing three axes of reference 
in a suitable manner, the co-ordinates of any particle are 
given by ja, kb.lc, where a, 6, c are three elementary lengths 
and J, k, I are integers at least, that is assumed to be the 
state of affairs at absolute zero of temperature with no 
thermal motion going on, each particle being held firmly to 
its equilibrium position by the forces arising from its neigh- 
bours. The particles are regarded as without structure ; 
so the model is " monatomic." If energy is given to the 
system, each particle will vibrate about its mean position, 
the displacements at any moment parallel to the axes being 
represented by , 77, , and the velocity-components by 
, 77, . If we make the well-known simple assumption that 
the elastic forces on a particle set up by the displacements 
are towards the equilibrium position and proportional to the 
displacement, the motion of each particle is simple harmonic. 
The energy is a quadratic function of the 3n co-ordinates 
(., 7] r , r ), and the 3n momenta (m ., mr} r , m r ), involving 
squares alone. The application of the statistical method 
follows the usual course, with a phase -diagram, in which are 
represented these 3n co-ordinates and 3n momenta, par- 

123 



124 STATISTICAL MECHANICS FOR STUDENTS 

titioned into phase-cells. Counting of complexions, deter- 
mination of the most probable state, etc., lead to the usual 
type of solution as to the number of representative points 
in a phase-cell, this number depending as ever on e~ Me 
where e is the energy corresponding to the centre of the 
phase-cell, and is, of course, the sum of kinetic and potential 
parts. The distribution constant /*, is still connected with 
the temperature by the relation kd = /x" 1 . To see this we 
can consider the solid immersed in a simple gas ; in the 
normal state they will have the same distribution-constant 
for the simple reason pointed out earlier in the treatment of 
mixtures of gases and of internal degrees of freedom in gas 
molecules. As energy is freely interchangeable between all 
molecules, gaseous or solid, there is only one variational 
equation for the energy, and thus only one multiplier, ^,, is 
involved for this equation in the solution by the method of 
indeterminate multipliers. By means of the pressure 
equation, we, as usual, identify the temperature of the gas 
as (i/x)*" 1 , and, of course, in equilibrium, the solid has the 
same temperature. 

A conclusion of great importance follows, as in Chapter V., 
viz., the equipartition on the average of the energy between 
the various components of displacement and of momenta, 
^ kO for each component, this being due to the absence of 
all but squared terms in the expression for the total energy. 
Thus it follows directly that the total energy is 6n times this 
elementary amount, or 3nk9. Hence the thermal capacity 
of the solid should be 3nk. For a gram-molecule, this is 
3JI where R is the gram-molecular gas-constant. This 
works out about 5*95 calories per degree. This well-known 
law, first pointed out by Dulong and Petit, is actually a 
good approximation to the truth for many monatomic 
solids, provided the temperature is sufficiently high, but the 
inference that the specific heat is independent of the tem- 
perature, is violently at variance with the facts. Just as in 
the case of diatomic gases there arises a serious discrepancy 
which classical statistical-mechanical theory has never been 
able to remove. The reader may think that our simple 
hypotheses are too restrictive ; but although with wider 



THE SOLID STATE 125 

conceptions as to the dependence of potential energy on 
displacement, we would obtain a different partition of the 
energy, the constancy of the specific heat as regards change 
of temperature would still emerge from the treatment, and 
this is an untenable conclusion, an asympotic fall of thermal 
capacity to zero as the temperature approaches absolute 
zero being one of most striking experimental facts dis- 
covered within the last twenty years. 

The removal of this discrepancy and the discovery of a 
satisfactory formula for the specific heat of a monatomic 
solid has, as stated, been effected by means of the quantum 
hypothesis. It is time to turn our attention to this way of 
escape from the various difficulties which have met us at 
several points on our way hither. 



CHAPTER XIV 

THE QUANTUM HYPOTHESIS 

14.1 The Three Stages in the History of the Quantum 
Theory. The first suggestion of the momentous change 
which has taken place in Theoretical Physics during the 
present century, was made in 1900, when Planck, in order 
to clear away a discrepancy between the experimental facts 
of black body radiation and the conclusions deduced from 
the current dynamical and electrodynamical theory, intro- 
duced the idea that the mechanism within an atom respon- 
sible for the emission and absorption of radiant energy, did 
not carry out this process in the continuous manner con- 
sistent with the laws of dynamics and of Maxwell's electro- 
magnetic theory, but in a discontinuous and " catastrophic " 
manner. It must be admitted that the reception of this 
notion was rather chilly ; in the mental atmosphere of that 
time anything which savoured of the " revolutionary " was 
frowned on ; Einstein had not as yet arrived. But he soon 
did ; in 1905, the very year which saw his first paper on the 
Relativity theory, he gave decided evidence of the fact that 
he always has been interested in other things in Physics 
besides the theory of space and time a fact not too widely 
known to the " popular " scientific public. He carried 
Planck's suggestion a step further and a very " shocking " 
step it was, even to those who were by that time prepared to 
listen to Planck. It introduced the idea of " atomicity " 
into radiation not merely in its moments of absorption and 
emission by matter, b.ut also in its propagation through 
space. Despite the fact that it ]gd to an immediate advance 
in the study of phenomena, such as fluorescence and the 
photoelectric effect, it was too much for the general scientific 
world, and it is only now just as the foundations of a real 
consistent Quantum theory have been laid down, that we 

126 



THE QUANTUM HYPOTHESIS 127 

can appreciate that Einstein's " light-quantum " idea was 
one of those flashes of insight vouchsafed now and then to 
the man of genius. Even so, Einstein did show two years 
later that Planck's suggestion in its original and less up- 
setting form could be applied to the elucidation of the 
theoretical difficulties which statistical mechanics encounters 
in dealing with specific heats. Planck, in addition, demon- 
strated a rather unexpected link between his hypothesis and 
the heat theorem of Nernst (the so-called third law of Thermo- 
dynamics), which was playing a great part in Physical 
Chemistry at this period. Planck, however, was no " revo- 
lutionary," and was also busy in recasting his original 
presentation of the quantum idea, so as to soften as far as 
possible the break with traditional conceptions. In 1912 
Debye and Born had subjected the whole problem of the 
specific heats of monatomic solids to a most searching 
mathematical analysis in the manner suggested by Einstein 
in 1907, and the result was a triumphant vindication of the 
power of this new weapon. The trouble concerning the 
specific heats of diatomic gases was also showing clear signs 
of yielding to the " new treatment." The leading physicists 
of the world were at last in their congresses and contributions 
to journals displaying the keenest interest in the " mystery 
of quantum." 

The second period in this eventful history was ushered in 
by three papers contributed by Bohr to the Philosophical 
Magazine in 1913. In them was first propounded the theory 
of the " Stationary States of an Atom," a theory which was 
the. clear descendant of Planck's first form of the, Quantum 
hypothesis and not the second. This period has been marked 
by the development of an uneasy partnership between the 
classical laws of dynamics and electrodynamics and two 
postulates of Bohr's, one a flat denial of a particular result 
in classical electron theory, the other, an ingenious modi- 
fication of Planck's law of emission for his " oscillator." 
The hostile partners have been driven in harness together, 
and " made to behave " by means of another ingenious 
notion of Bohr's, the " Principle of Correspondence." This 
patchwork affair has, despite the incongruity of the situation, 



128 STATISTICAL MECHANICS FOR STUDENTS 

been at once incentive and guide during a dozen years 
amazingly fertile in theoretical investigation and experi- 
mental research. The debt which experimental physics owes 
to this " Classical-Quantum theory" is large beyond question; 
yet equally certain is it that this was not a " theory " in the 
accepted sense of a perfectly consistent body of principles 
from which all the essential experimental facts could be 
deduced. This period closed in the autumn of 1925, when 
Heisenberg, in a paper which will probably rank with 
Einstein's first relativity paper of 1905, as an epoch-making 
communication, pointed clearly the direction in which, we 
had to go for a genuine escape from all our theoretical doubts 
and misgivings. Then began the third period, a period of 
the construction of a genuine Quantum theory of atomic 
phenomena. Already in four years physicists can feel that 
their science is based once more on a foundation which, 
although not as yet complete, is as far as it goes firm and 
self -consistent. To the philosophical mind it is of interest to 
observe one common feature of Heisenberg's contribution 
to the theory of the microcosm which we call the atom and 
Einstein's contribution to the theory of the large-scale pheno- 
mena of the " world." Einstein pointed out that we were 
unduly hampering our ideas by trying to reconcile the 
independence of the velocity of light with respect to the 
frame of observation and the assumption of an absolute 
space ; as the former is a physical fact, he suggested that 
the latter should be abandoned and the mathematical 
treatment suitably modified, and proceeded to show how 
it could be done. Heisenberg also indicated the restrictive 
effect on the development of atomic theory of the endeavour 
to run together the idea of electron motion within an atom, 
subject to the usual kinematic ideas and regarded as a 
resultant of a simultaneously existing group of harmonically 
related components, with the physical fact of spectroscopic 
lines subject to a law of frequency so different to the law 
of a series of harmonic terms ; so he surmised it would 
be wiser to cease to trouble ourselves about electron move- 
ments which never come into the actual field of obser- 
vation. No doubt this would require a recasting of the 



THE QUANTUM HYPOTHESIS 129 

mathematical treatment and Heisenberg gave an indica- 
tion of the way to proceed. Another feature of interest 
to the scientific historian is the fact that in both instances 
the pure mathematicians of the nineteenth century had 
actually, without any prevision of the use the physicists 
were destined to make of them, invented the suitable mathe- 
matical conceptions and developed and perfected the 
necessary technique ; the calculus of tensors and the calculus 
of matrices were both to hand when the right moment for 
their physical applications arrived, although to be strictly 
accurate in our statements, both Einstein and Heisenberg had 
to be informed of their good fortune by the mathematicians. 
Naturally our subject, Statistical Mechanics, is being 
brought into line with the Quantum Mechanics of this new 
period. Just as naturally that is a matter outside the pro- 
vince of a book of this nature.* The author, however, begs his 
youthful readers not to be too downcast on that account. 
The kind of feeling that " classical and classical-quantum 
stuff " is dead and done with, and that it is just so much 
waste time to bother about it is very unjustified. We travel 
quickly in these days no doubt, but I doubt if any serious 
teacher of physical science can see how it would be possible 
to introduce the immature and youthful mind to the newer 
knowledge without an adequate training in the traditional 
conceptions, the manner in which they synthesised the older 
knowledge, and the manner in which they were modified and 
replaced by broader ideas. After all it was to deal with 
difficulties in statistical mechanics that " Quantum " was 
first invented, and nearly all the work of the first period was 
concerned with this purpose ; and as regards the second 
period, the conception of " stationary states " is still required 
in order to follow the generalisation to the new formula- 
tion of Quantum theory. As far as the matter of " seeing 
results as quickly as possible " is concerned (which we at the 
outset assumed to be the desire of the majority of readers), 
there is no need to be alarmed ; the first and second periods 
have provided them in abundance. 

* Nevertheless, an Appendix at the end of the book will give the reader 
some idea of what has happened very recently. 



130 STATISTICAL MECHANICS FOR STUDENTS 

14 . 2 Planck's Constant. At the time of Planck's first 
suggestion with its flavour of " heterodoxy," one of the 
problems agitating the minds of the physicists was con- 
nected with the discrepancy between the facts of black body 
radiation as discovered by an improved technique in radio- 
metric measurements and the theoretical laws as deduced 
from dynamical principles and the equations of the electro- 
magnetic field. Thermodynamical reasoning carried the 
work far enough to recognise that the formula for the 
density of radiation in a uniform temperature enclosure has 
a certain general character ; * in it, however, there occurs 
an unknown functional form, unknown, i.e., in the sense that 
thermodynamics alone can give us no information about it. 
For progress towards its discovery an appeal to electro- 
magnetic and dynamical theory had to be made. Un- 
fortunately, if that appeal was made in a strictly " lawful " 
way, the result was seriously at variance with the facts. 
Planck working hard at the problem on its theoretical side, 
gradually narrowed down the region in which the fallacious 
step was to be found. At last he put his finger on it ; it 
turned out to be the assumption of equipartition of energy 
on the average between co-ordinates and momenta in any 
molecular system which contribute squared terms to the expres- 
sion for the energy. The italicised words are important. No 
doubt there is no equipartition if that condition is not satis- 
fied, but Planck was using as a model radiating and absorbing 
mechanism the harmonic oscillator (the faithful ally of the 
mathematical physicist, which had never yet failed him), 
and its energy was a sum of squared terms. It availed nothing 
to point out that after all this was a very crude model of the 
radiating processes in an atom. The nature of the reasoning 
was such that any conceivable mechanism following dynamical 
laws should yield the proper result. But the law of equi- 
partition of energy is derived from statistical-mechanical 
reasoning, and that is how our subject became mixed up 
with all the trials and tribulations of that period. 

To appreciate Planck's hint as to how to escape from the 

* Wien's Displacement Law. 



THE QUANTUM HYPOTHESIS 131 

dilemma, we will introduce the reader to his way of deriving 
the law of equipartition. His statistical reasoning in his 
first papers did not follow quite the same course as that 
employed in our earlier chapters, but as a matter of fact, his 
particular way of choosing complexions and counting them, 
has reappeared quite lately in the literature of the subject. 
Added to that, it has an interest of its own, and is, of course, 
quite sound ; so it should be known to any one interested 
in the elements of the subject. 

Planck conceived a given number of oscillators vibrating 
about fixed mean positions in an all-pervading reservoir of 
energy, viz., the field of full radiation. Between the field 
and the oscillators there was flux and reflux of energy. Just 
as in the beginning of our statistical reasoning we had to 
postulate finite phase-cells, which we afterwards reduced to 
mathematical infinitesimals, so he postulated finite elements 
of energy, each passing as it were, entire, and not con- 
tinuously in small infinitesimal elements, at exchange 
between oscillator and field. Call each element 77. Suppose 
that at any moment the n oscillators have between them c of 
these elements, so that their total energy is cr). How many 
ways can this be done ? All the c elements might be in the 
first oscillator ; represent this symbolically by af, or all in 
the second represented by a/, and so on ; or c 1 of them 
might be in the r th and one in the s th represented by a r c ~ l a 8) 
and so on. As all the elements of energy are supposed to be 
indistinguishable the number of ways of partitioning the 
energy E among the n oscillators is the number of terms in 
the expansion * ^ 



and this is known to be 

w -1 C - I 

(*+*-!)! 
*!(*-!)! ' ' ' ' (14 ' 2 ' 1 > 

>o<Y- 0! ; 

Now when in equilibrium with the radiation, the oscil- 
lators will have the temperature 6 of the radiation, and 
possess a definite energy E. The whole system radiation 
and oscillators will then statistically be in its most probable 

E 2 



132 STATISTICAL MECHANICS FOR STUDENTS 

state. If there are N elements of energy altogether present, 
the total number of complexions in this state is 

/(N-c) (" + -*)' . . . (14 . 2 . 2) 
nl (c 1)1 

where f(x) is the number of ways of distributing x elements 
of energy in the radiation, and c E/T? ; for with any one 
way of distributing the N c elements in the temperature- 
enclosure, there can be combined one way of distributing 
c elements among the oscillators to yield one way of distri- 
buting all the N elements among the various parts of the 
whole system. As usual, we take the entropy of the system 
in this equilibrium state to be k times the logarithm of 
(14 . 2 . 2), together with a constant term which for our 
purpose may be ignored as it will disappear in the differentia- 
tion to be carried out presently. It is plausible to regard 
the two parts into which this expression falls as the entropy 
of the radiation and the entropy of the oscillators respec- 
tively. So if S is the entropy of the oscillators 

" 



== Jclog 



n\(c \)\ 

(n + c)\ 



n ! c I 

= k (n + c) log (n -{- c) k c log c k n log n. 

If the temperature of the enclosure be altered to 8 -f 80, 
the 'energy of the oscillator system will be altered to E -f 8E 
and its entropy to S + SS, an d we know from thermo- 
dynamical reasoning that as no external work is done 

j _ ! 

"SE "~ J 

in the limit. But SE ^8c. Hence 
l^J^S 

~~ 77 dc ' 



= - (1 + log n + c 1 log c) 

V 

* We are assuming that the elements of energy are much more numerous 
than the oscillators so that (n + c)/c is practically unity. 



THE QUANTUM HYPOTHESIS 133 

k , n + c 
= -log ! . 

i c 

Thus - = <? 1 - 1, 
c 

or ?! = _J? (14.2.2) 

^ gj//Jt0 j v ' 

This reasoning may seem extremely abstract to the 
reader, who may be pardoned if he feels that the radiation 
is a very shadowy kind of material to partition elements of 
energy among ; nothing so tangible for example, as the 
"solid" particles of our earlier "games of chance." It 
would not be so to any one imbued with present-day ideas 
of the essential " substantiality " of the radiation, in so far 
as it possesses all the so-called mechanical properties of 
matter, mass and momentum, as well as energy. Be that as 
it may, he may feel reassured when he learns that the result 
(14 . 2 . 2) for the average energy of an oscillator is equi- 
partition of energy, if we just carry the analogy with our 
former procedure to its logical limit, viz., assume that the 
elements r\ are mathematical infinitesimals, as we formerly 
assumed the phase-cells, in terms of which we defined com- 
plexions and states, to be infinitesimals in our final calcu- 
lations. If we do so, the right-hand side of (14 . 2 . 2) is equal 
to 



which approaches the value kO as t\ approaches zero. And so 
the average energy of the oscillator becomes kd, and this 
agrees with our former result, | kO of kinetic, and J k0 of 
potential on the average. 

However, it was the really brilliant idea of Planck just to 
refuse to go to the limit. With a flash of insight he saw that 
in refusing to do so, and in thus denying the equipartition 
law (as we shall see presently) lay salvation. Here We must 
take on faith a result which concerns the electrodynamical 
side of Planck's complete argument as distinct from the 
purely statistical with which we have been immediately 



134 STATISTICAL MECHANICS FOE STUDENTS 

concerned. In that he had deduced the relation between 
the energy-density of the radiation and the average energy 
of an oscillator in temperature equilibrium with it. If he 
took the latter to be k6, the result was at variance with 
experiment ; so he took (14 . 2 . 2) as it stands with T? finite, 
and considered the conclusion to be drawn from that. At 
once he saw that in order to satisfy the general character of 
the formula for the energy-density, derived from purely 
thermodynamical reasoning and referred to on page 130, 
77 would have to be proportional to the frequency of the 
oscillator and the constant of proportionality would, more- 
over, be a universal constant. So calling the frequency v he 
wrote 

7] = JlV, 

and so 

5= ^ . . . (14.2.3) 

n exp (hv/k0) I 

From this could be derived at once a formula for the energy 
of full radiation as distributed among its various frequencies. 
Within a short time the experimental physicists were con- 
vinced that it was correct, and derived from the observations 
a value for h which has since been confirmed in several 
experimental researches inspired by other applications of 
the quantum idea. The accepted value for h is 

6-55 x 10~ 27 erg-seconds, 

for, as will appear at once, its physical dimensions are the 
product of energy and time, i.e., the dimensions of action. 

14 . 3 The " Quantised Paths " of an Oscillator. We have 
given Planck's original treatment (somewhat amplified) of 
the statistical side of the problem. It will be interesting to 
deal with the matter in a manner more in keeping with the 
methods which we have used hitherto, now that we appre- 
ciate Planck's break ,with tradition. This will have the 
added advantage that it will give us also a truer idea of 
where the " quantisation " is really situated. In terms of 
the conceptions of section (14.2) we naturally speak of 
" quanta of energy/' but if there is any idea latent in the 
reader's mind that this must imply discrete " atoms " of 



THE QUANTUM HYPOTHESIS 135 

energy preserving an identity through all vicissitudes, as 
we have been accustomed to assume for atoms of matter, 
he must disabuse his mind at once of this idea. Planck 
himself would have none of it in those days and when 
Einstein, in his study of the photoelectric effect, propounded 
quantum views savouring very much of such heretical 
notions, Planck protested vigorously, and recast his whole 
presentation in such a manner as to bar out this idea, which 
was really at that time repugnant to the sense of con- 
tinuity produced by the theory and the experimental facts 
of the propagation of radiation. No doubt the oscillator 
could only take in " lumps " of energy, hv, or get rid of them 
if it emitted or absorbed at all. The reader will realise that 
the whole of the argument would be upset, if there could be 
fractions of 77 in an oscillator as well as integral multiples. 
But once out, the energy merged into the continuous field. 
Nevertheless, one conclusion could not be avoided ; between 
emissions and absorptions the vibrations of the oscillator 
could only be executed with one of a discrete series of 
amplitudes ; i.e., it could only exist in one of a discrete set 
of " quantum states." We are familiar in ordinary mechani- 
cal reasoning with the notion that a harmonic vibrator can 
be given any of the infinite number of amplitudes between 
zero and some upper limit. But that notion will not serve 
here. We can very readily select those " quantised " ampli- 
tudes, by means of the information already obtained. 

In the symbolism of section (5.1) the energy of the oscil- 
lator is 



. 

or + 

2a 2 

If this has a constant value, e, the representative point of 
the oscillator in a phase-diagram will be on the ellipse 



2 6/6 2 e a 

Its semi-axes are (2 /&)* and (2 ea)*, and its area is the 
product of these by TT, i.e., 2 TT e (a/6)*. 



136 STATISTICAL MECHANICS FOR STUDENTS 

But (6/a)* is the pulsance, 2 n v, of the vibration, and so the 
area of the ellipse is 

- (14.3.1) 

V 

In consequence, if in this quantum state the oscillator holds 
r " quanta of energy of frequency v," its elliptical phase- 
path has the area 

rh (14.3.2) 

Thus the discrete quantum paths of the representative point 
in the phase-diagram of the oscillator are a series of similar 
ellipses separated from each other by elliptic annuli of area 
h, the lowest quantum condition being represented by the 
origin. Thus no matter what the value of v is, i.e., no matter 
what is the value of the quantum of energy, the phase- 
diagram is divided up into areas always of the same magni- 
tude. The only quantum which hab so far any claim to 
atomicity is thus the " quantum of action," for it is action 
which is represented by an area in the phase-diagram. It 
will be wise to remember that a quantum state is not a 
static condition ; the oscillator vibrates to and fro, and the 
representative point keeps rushing round the appropriate 
ellipse between the ' c catastrophic ' ' emissions and absorptions . 
The statistical problem is now easily dealt with in our 
more customary manner. A complexion of the system of 
oscillators is determined by the manner in which we assign 
individual, identifiable points to the various ellipses ; there 
will be some outer limit settled by the whole energy of the 
system. The ellipses take the place of the phase-cells in our 
previous arguments; an interchange of points between 
ellipses alters the complexion, but not the statistical state ; * 
a mere shifting of points along the ellipses does neither. 
The number of ways of assigning n points to the origin, % 
to the first ellipse, n 2 to the second, and so on, is 

nl 



n. n. n 9 



* Do not confuse the two uses of the word " state " from this point 
onward. We speak of a " quantum state " of an oscillator, and, later, of 
an atom ; " Statistical state " still refers to the whole system of oscillators 
or atoms. 



THE QUANTUM HYPOTHESIS 137 

We then proceed as before. The logarithm of this is varied 
and the variation put equal to zero. The conditions for con- 
stant total number and total energy introduced, and we 
arrive at the usual result. In the most probable state the 
number of oscillators in the r** quantum state is 

C e" w 
where r = rhv. 

The constant //, is, as usual, identified with (k6)~ l ; for we 
are really considering the oscillators now as denizens of gas 
atoms in the manner of Chapter V. 

To determine C, we have 

C (1 + e~^ + e~^ hv + e~ 3 ^ + ...... ) = n* 

or C = n (I e-"*") . . . . (14.3.3) 

The whole energy in the oscillators is 

C 2 T e-*"' 

ro 

which is equal to 

C hv (e~* hv + 2 e~ 2 * hv + 3 e~* v + ...... ) 

= rihv (1 e-* hv ) e'"** (1+2 e~^ hv + 3 e~^ hv + ...... ) 



nhv 



e hv I 

or the average energy of an oscillator is 

hv 

as before. 

It will be seen that in the most probable statistical state of 
the system, the majority of the oscillators are in the lowest 
quantum state represented by the origin, the number in the 
next quantum state is obtained by multiplying the former 

* The unity in the series arises from the lowest quantum state, r = o. 



138 STATISTICAL MECHANICS FOR STUDENTS 

by exp ( hv/kd), in the next by exp ( 2 hv/k6), and so 
on. If the temperature 6 decreases the multiplier e""^", 
becomes smaller and smaller approaching zero as 6 approaches 
zero. Thus as the temperature falls, the oscillators tend to 
crowd into the state of zero energy. If 0, on the other hand, 
increases, the exponential factor increases gradually having 
unity as its limit when 6 is infinite. Thus there is with 
rising temperature a tendency towards more equal distri- 
bution of the oscillators among the quantum states. 

14.4 Planck's Alternative to the Strict Conception of 
Quantum States. It is possible that the reader may not 
quite realise the point of the extreme hostility to Planck's 
views at the time. The assumption of quantum states seems 
harmless enough in its way, and the statistical argument, 
at all events in the form outlined in section (14.3), appears 
as unimpeachable as it was in earlier applications. That is 
true enough ; it was not the statistical part of his investiga- 
tion which failed to command general assent at first. But 
there was another part just as necessary to the argument, 
and in that part Planck assumed that the amplitude of the 
oscillator could vary in a continuous manner. The oscil- 
lator is subject to the electromagnetic forces in the radiation 
and those forces which have the same frequency as the 
oscillator, or frequencies relatively near, impress an oscil- 
lation of increasing amplitude on it an illustration of the 
well-known phenomenon of resonance. The safeguard 
against undue heaping up of energy in the oscillator is its 
own radiation due to the accelerated motion which its 
vibration implies. The balance between absorption and 
radiation is required to keep it in an average condition as 
regards energy, whether that condition be equipartition or 
any other. The mathematical equation which expresses 
that balance leads to the relation between the density of 
each constituent of full radiation and the average energy of 
oscillators with the corresponding frequency. Thus it will 
be seen that this side of the argument was absolutely 
essential to Planck's final result, and in working it out he 
had to assume the continuous increase or decrease of the 
energy in the oscillator. In an endeavour to minimise the 



THE QUANTUM HYPOTHESIS 139 

contradiction as far as possible, and as a protest against what 
he considered to be an unnecessary extension of his idea, he 
recast the statistical argument so as to suit at all events the 
assumption of continuous absorption, although he still had 
to keep discontinuous emission in it. Though it has not 
much place in the general exposition of quantum physics, 
as it developed later, it should be known to the student of 
these matters, as it suggests one conclusion which may 
possibly be justified soon as an experimental fact. To 
appreciate its point of difference with the first method, let 
us conceive that the phase-diagram is partitioned into phase- 
cells by drawing equal energy ellipses very near to one 
another. The phase-cells are the elliptic annuli, each of 
area Se/v where Se is the step of energy between one ellipse 
and the next. By classical methods the density of points 
in an annulus is proportional to e~^ and the number in it 
is D e~^ e Se/v where P is a constant given by 

Dr 

__ e-^de =n . . . . (14.4.1) 

VJ 

so that D = Ufjiv. In obtaining this, we have, as formerly, 
narrowed down the cells to infinitesimal dimensions in the 
last resort. In the equilibrium state as many representative 
points pass outward in a given time across any given ellipse 
(absorption) as pass inward across it (emission), and an 
inward, or outward journey could start from any point. 
But in Planck's second form of his theory, an inward journey 
could not start from any point ; once a representative point 
has passed out beyond one of the critical ellipses of area 
A, 2 A, 3 A, etc., it cannot move inward ; not until it has 
reached the next critical ellipse is that possible. In short, 
emission is only possible at certain critical amplitudes. At 
such critical moments of its history the oscillator may 
radiate or it may pass on into the next " zone of safety " 
free to gather up energy for another spell without danger of 
loss. But some of the oscillators are bound to radiate when 
they reach any critical ellipse ; otherwise the balance 
between absorption and emission over relatively long periods 
would not be maintained. When one speaks of an " outward 



140 STATISTICAL MECHANICS FOR STUDENTS 

journey/' one visualises the representative point on the 
phase diagram executing a spiral path gradually widening 
out from the origin ; such a path may terminate suddenly 
at the first critical ellipse and restart at once from the 
origin, or it may continue widening out uniformly in the 
annulus between the first and second ellipses until it reaches 
the latter, when it may still continue or suddenly stop and 
make a fresh start from the origin ; and so on. Thus the 
number of points in any h annulus is at any moment less than 
in the one just inside it and more than in the one just out- 
side ; but just how much less and how much more ? Well, 
one can, as before, consider these h annuli as finite cells, and 
assign n l points to the first, n 2 to the second, and so on, and 
proceed to work out the most probable state ; but in writing 
down the constant energy condition we realise that all the 
points in one of these cells do not correspond to the same 
energy ; we cannot assume that, as we are not ultimately 
going to make these cells infinitesimal in size. A plausible 
proposition is that the points in an h annulus are uniformly 
distributed over it, and in that case when we write the 
equation 



e r is the mean energy of the r th annulus, i.e., it is equal to 
(r ) hv. The usual result emerges 

n r = C e-^ r 
where 

<r = (' - i) *" 
As usual C is determined by 

C e-^r = n, 

r=l 

or 

C e-f (1 + e-* + e~ 2x + ...... ) = n, 

where x is for the moment written instead of phv. 
Thus 

C = n e 2 (1 O- 



THE QUANTUM HYPOTHESIS 141 

The whole energy E is given by 
E = C 2 r e-"*' 



3 e~ 



= ~ 

+ 2 (e~ x + 2e- 2 * + 3 e~ 3x 

nhv . , ., ~ x 

= - + nhv (1 



2 ' v ' (1 - 



_ 

___ j 



2 e* - I 
Thus the average energy of an oscillator is now 

hv + hv n4 4 2) 

exp (hv/k0) 12'''* 

There is an interesting difference between this result and 
(14.2.3). If we make 6 gradually approach zero, the expres- 
sion (14 . 4 . 2) does not approach zero but hv/2 as its limits ; 
thus at absolute zero the system of oscillators would have an 
energy of amount nhv/2. Planck could still make this new 
expression yield the satisfactory expression for black body 
radiation, and he was also able to use the new model in an 
explanation of the (then) puzzling features of the photo- 
electric effect without recourse to Einstein's light-quanta. 
This being so, he felt justified in asserting that here there 
was theoretical evidence of the possibility that at absolute 
zero of temperature matter is not quite devoid of energy. 
The existence of " nul-point energy " is suggested. 

The number of oscillators in one of the finite energy ranges 
defined by the critical states of emission bears to the number 
in the range just below the ratio eT x to unity. Thus the 
chance that an oscillator will radiate when it reaches a 
critical state is 1 e~ x , provided it radiates the whole 
amount which it contains and restarts a fresh accumulation 
of energy from that condition. In his statistical-electro- 
dynamic argument, Planck assumes that emission has this 



142 STATISTICAL MECHANICS FOR STUDENTS 

character, and introduces a definite postulate about the 
chance of emission in a critical state ; these hypotheses 
correspond to the assumption made above in this simpli- 
fication of his argument, that the points in an h annulus are 
equally distributed. The reader will observe that in his 
first exposition emission had not of necessity this character ; 
the representative point could jump about from ellipse to 
ellipse, in its changes of state, inward or outward ; if inward, 
the leap was not necessarily right to the origin. 



CHAPTER XV 

THE THEORY OF THE STATIONARY STATES OF AN ATOM 

15.1 The " Action-Integrals " of a System. Notwith- 
standing Planck's conservative tendencies the movement 
which he initiated took his first idea as its guide for further 
development. To see how this came to pass, we must deal 
with some general dynamical features of more complex 
systems than a simple harmonic oscillator ; it is too crude 
a model for further exposition. For one thing the assumption 
that even for an oscillator with but one degree of freedom 
the period is independent of the amplitude, is not really 
true of the many physical systems which the oscillator is 
taken to represent in elementary statements the pendulum 
for instance, or the body bobbing up and down at the end of 
a spring. But if this is so, the apparent simplicity of 
quantising by means of the energy content disappears, as 
that clearly depends on a constant frequency. But we saw 
that another way of declaring how to select the quantum 
states of the harmonic oscillator is to say that the choice 
will fall upon those amplitudes for which the accumulated 
action in one period is once, twice, thrice, etc. the value h. 
This choice is still open to us, and during the second period, 
referred to in section 14 . 1, it was the keynote to the whole 
situation. 

The restoring force which controls the oscillator not being 
proportional to displacement beyond minute values of the 
latter, it turns out, as stated, that the frequency of the 
oscillation depends on the amplitude. The oscillator will, 
however, in any actual vibration, have a definite momentum, 
p, for a given value of displacement, q. (It must be clearly 
understood that " damping " is excluded from consideration.) 
The integral of pdq can be calculated from the knowledge of 
the particular law of force, and if this integral is calculated 

143 



144 STATISTICAL MECHANICS FOR STUDENTS 

for the complete period which exists for this given vibration, 
we shall denote its value by J. This accumulated action in 
one period is usually termed the " action-integral " of the 
system. It is like an integration constant occurring in the 
statement of q as a function of the time ; it is a " constant " 
during any movement involving a given amplitude, but 
varies in value if the swings are altered so as to involve a 
wider or narrower range. Amplitude, frequency and energy 
can all be expressed as functions of J. It will be recalled that 
in the case of a harmonic oscillator, if the energy of a par- 
ticular vibration is e, then the accumulated action in one 
period is e/v, so that the quotient of the former by the latter 
is the frequency, or to put it another way, any finite change 
in the energy if divided by the change in the action-integral 
accompanying it, gives the frequency of oscillation. The 
result for an " anharmonic " oscillator is a generalised form 
of this. Let E(J) be the function of J, which is equal to the 
energy ; E(J + SJ) is equal to the energy for a slightly 
different amplitude with an action variable J + 8 J. It can 
be proved that on dividing E(J + SJ) E(J) by SJ and 
going to the limit, we obtain the frequency of the oscillation 
with the action-integral J. In short, the frequency of 
oscillation is d^(J)/dJ. In addition to this modification of 
the results it must be borne in mind as well that we cannot 
write 

q = A cos (cot e) 
where 

*E(J) 
W - 2 "-d3- 

The oscillation is not a simple harmonic one. Instead it is 
known that the correct result is 

q = Aj, cos (cot <j) -f A 2 cos (2 cot < 3 ) + A 3 cos 
(3 oot - < 3 ) + ...... 

where ^ 1? < 2 , </>& the epoch-angles, and A lf A 2 , A 3 , 

are known functions of a second integration constant 

and J.* The selection of the quantised oscillations for this 

* The equation of motion is of the second degree, and in its complete 
solution two integration constants must make their appearance. 



STATIONARY STATES OF AN ATOM 145 

oscillator is, if we adopt the method used for the model 
atoms of Bohr's theory, made by choosing those whose 
action-integrals are whole-number multiples of h. To be 
strictly accurate it was discovered as time went on that 
some experimental results were more easily brought within 
the bounds of the theory if this selection were interpreted a 
little more loosely and " half -quantum numbers " allowed, 
so that the action-integrals might be put equal to rJ or 
(r + |)J where r is a positive integer. As a matter of 
interest this extension of the quantising principle does for 
certain systems emerge quite naturally from the most recent 
formulation of Quantum mechanics. It should, however, be 
noted carefully that this extension does not bring in its wake 
any necessity for " half -quanta of energy." The reason why 
will appear presently. 

To go on to a system with two degrees of freedom, such as 
the ordinary pendulum might be considered to be, an 
interesting feature of the vibrations of such a system can 
be observed very readily by giving a pendulum a slightly 
elliptical swing. It soon appears that the orbit of the bob 
is not an ellipse, nor indeed any closed oval curve. Roughly 
we say that the plane of vibration of the string is rotating 
round. Strictly, of course, there is no such plane. Another 
mode of expression is to say that the bob is going round an 
elongated ellipse, whose long axis is slowly " precessing " in 
a horizontal plane in the same sense as the motion of the 
bob round the orbit. What is really happening is that the 
bob does not return to the same place at its places of greatest 
and least distance from the centre. This arises from the 
fact that it has really two distinct periods of vibration, which 
may be nearly equal but are not exactly so. The amplitude 
one way is quite large ; at right angles to that way quite 
small. It will happen, of course, that if the two periods have 
values which are commensurable then ultimately the curve 
after many convolutions will " re-enter " into " itself," and 
the body proceed to execute the same path once more. But 
if the periods are incommensurable, this is not so. The 
simplest way of bringing out the two periods is to realise 
that there is a certain interval during which the radius 



146 STATISTICAL MECHANICS FOR STUDENTS 

vector from the centre changes from its greatest value to its 
least, then to its greatest, once more to its least, and finally 
to its greatest. This is one period, the period of " libration " 
of its radius vector. (We are really at the moment thinking of 
a vibration truly in a horizontal plane; actually the bob of a 
pendulum travels on a spherical surface.) On the other hand, 
the angle which this radius vector sweeps out in the horizontal 
plane, increases by 2 77 in a different time ; in this time cos 
and sin 0, where is this " azimuthal " angle, librate in value 
between the maximum and minimum values, + 1 an< l 1? 
and back again. In this system the accumulation of the 
action goes on in two ways. The vibrating body has a 
certain angular momentum round the origin, mr*8, where r 
is the radius-vector to its position on the oval path ; if we 
integrate mr 2 9d0 between and 2??, we obtain one action- 
integral J 1( The body also has a linear momentum m'r to 
and from the origin ; if we integrate m'rdr throughout the 
period of radial libration, we obtain a second action-integral 
J 2 . The energy in any prescribed orbit can be found from 
a knowledge of its two action-integrals ; if E(J 15 J 2 ) is the 
function which is equal to the energy, then the two fre- 
quencies azimuthal and radial are 



1? J 2 ) , SEfa, J 8 ) 



, 
d 



X 9J 2 

respectively. That is known from dynamical theory. The 
expressions for r and in terms of t take the form 



oc oo 



r == 2 H A r8 cos (r^ + sa> 2 ) t + <j> rs 



:== > 2j B cos ] (I*CL)I -f- 5cL)o) t -f- <p o - . 

r.-r= oo = oo I ' 

Here the A r5 , B r ^ and ^ are functions of J 1} J 2 and two 
other integration constants, and as stated, aj l = 2-rr SE/SJj, 

C0 2 = 277 3E/9J 2 . 

We have here an example of a " conditionally -periodic " 
system. The system may not apparently be periodic ; for 
we have seen that if o^ and co 2 are incommensurable, the 



STATIONARY STATES OF AN ATOM 147 

vibrating body never really returns to any former position, 
but the solution shows how the periodicity is latent in 
artificially separated 'parts of the motion. 

The suggestions for quantising in this case follow the same 
lines as before ; the quantised paths are chosen so that J l = 
r-Ji and J 2 ~ rji where r l and r 2 are positive integers (or 
numbers such as r + ^ may be involved). 

15 . 2 Bohr's Postulates for the Atom. Possibly the 
reader may now begin to have some idea of the quantisation 
of orbits, and how action and Planck's h constant are in- 
volved in it. Taking the hydrogen atom with its single 
electron, it has to be observed that the strictly elliptic orbit 
of the electron round the nucleus as a focus, deduced from 
the simple inverse square law of attraction, is too ideal. 
Owing to a variety of causes, the forces caused by neighbour- 
ing atoms, or external fields imposed by an experimenter in 
the study of the Zeeman and Stark effects, or the relativity 
change of mass in the electron due to changing speed in the 
orbit, the orbit is not really re-entrant. Just as we have 
pointed out in the case of the two-dimensional anharmonic 
oscillator, there are two periods, one involved in the libration 
of the radius vector between its extreme values, one in the 
increase of the azimuthal angle by 2??, and a repetition in the 
values of its circular functions. Two action-integrals are 
involved just as before, and one of Bohr's fundamental postu- 
lates consisted in assuming that, despite the deductions of 
classical dynamics, the orbits whose action-integrals are equal 
to integral multiples of h have an inherent stability, non- 
dynamical to be sure, but physical in the sense that the 
assumption could be used with great effect to unravel for 
the first time some of the intricacies of spectroscopic obser- 
vations which had hitherto baffled physicists. The remaining 
non-quantised paths, just as dynamical and from the point 
of view of Dynamics just as "unstable" as the quantised, 
are irrelevant to the explanation of the physical facts. 
Conceivably, if undisturbed, the electron could remain in a 
quantised orbit for ever, but if disturbed, it would have to 
find a new semi-permanent home in another quantised orbit, 
not in any of the mechanically possible, but " quantically " 

L 2 



148 STATISTICAL MECHANICS FOR STUDENTS 

impossible orbits. The atom has a discrete number of 
"stationary states/** It should be mentioned that 
Sommerfield and Schwarzschild gave considerable assistance 
at the outset on the mathematical side in applying the 
quantising rules to particular problems. 

The second postulate of Bohr concerns the " quantum- 
jumps " during which the atom leaves one stationary state 
and enters another. In Planck's early oscillator theory the 
quantum of energy is determined by the frequency of the 
oscillator vibration ; h if multiplied by the latter gives the 
former. In view of the multiplicity of kinematic frequencies 
now involved, no success in that direction seems possible. 
Bohr's ingenious modification lay in dividing the change of 
energy between two states by h to obtain the frequency of 
the emitted or absorbed radiation. Optical frequencies 
ceased to be identified with kinematic. Terra firma began 
to appear in the region of spectroscopy where, in the word 
of the late Lord Rayleigh, there had been formerly a " bog." 

We have assumed that the orbit of the electron is in one 
plane ; virtually that limits the degrees of freedom to two ; 
two co-ordinates are sufficient to determine the position of 
the electron ; there are two distinct action-integrals for any 
orbit, classical or quantum, and in consequence two funda- 
mental frequencies involved in each orbit. But the degrees 
of freedom are really three ; our limitation of the orbit to 
one plane has concealed that fact. But if we apply some 
external force to the electron, such as a magnetic field, the 
normal to the plane of the orbit precesses round a line 
through the nucleus parallel to the field, just as the axis of 
a spinning top precesses round the vertical. To define the 
electron completely now we must have a co-ordinate 
defining the plane of the orbit or the normal to it, in addition 
to the radius vector and azimuth of the electron in the orbit. 
This can be chosen to be the angle made by a plane contain- 
ing the normal to the orbital plane and the field line through 
the nucleus with any reference plane containing the latter 
line. This precession involves additional energy ; and also 

* Note that *' stationary " does not mean " static." There is plenty of 
movement in a stationary state. 



STATIONARY STATES OF AN ATOM 149 

a new angular momentum round the field line. Prom the 
latter, using dynamical laws, can be calculated a third 
action-integral equal to the amount of action accumulated 
in this fashion in the period of precession, the third period 
of the system. Classically this may have any value ; the 
requirements of the Quantum hypothesis limit it again to 
integral multiples of h. This amounts to stating that the 
orbit can only have a set of discrete orientations with respect 
to the external influence. Hence arises " space-quantisa- 
tion," and a third quantum number enters in the selection 
of stationary states. 

Nor is this all. In our solar system the planets spin on 
their axes as well as rotating round the sun in their orbits. 
So in recent years the " spinning electron " has come along 
to give us a hand in this entertaining puzzle of defining 
stationary states so as to conform to the spectroscopic, the 
magneto-optic, the electro-optic and the thermal evidence. 
This involves another angular momentum and period ; 
another action-integral and a fourth quantum number. 

All this has been written in connection with the single- 
electron atom. As a matter of fact, in view of a special 
feature of Bohr's theory of optical spectra, it holds good in 
an approximate way for much more complicated atoms. 
Further we can hardly go without transcending the limits 
of space and possibly the reader's ability to follow the 
necessary analytical statements. In the last resort, if there 
are I electrons, we can (regarding the nucleus as sufficiently 
massive) think of the system as having 31 degrees of freedom. 
The radial vectors of the electrons will librate between 
maximum and minimum values, the angles necessary for 
co-ordination will also librate or their circular functions will. 
These are necessary conditions of periodicity and stability, 
even if they are not sufficient. The whole theory can be 
worked out under certain definite mathematical assump- 
tions. There are 31 action-integrals involved, and 3Z fre- 
quencies given by the same rule as before 



150 STATISTICAL MECHANICS FOR STUDENTS 

Quantisation chooses those states of orbital motion in which 
the J r are whole-number multiples of k, and so each station- 
ary state corresponds to a definite set of 31 integers. In fact, 
in recent theory less and less interest is being shown in the 
attempt to make pictures of the orbits or give analytical 
expressions for them. The state is determined by its quan- 
tum numbers. If we know the energy as a function of the 
quantum numbers, i.e., of the action-integrals, the change 
of energy in a jump from one state to another can be cal- 
culated, and the frequency of the emitted radiation obtained 
by Bohr's second postulate. The results must, of course, 
agree with spectroscopic evidence ; that acts as a check on 
any hypotheses we introduce for formulating the energy in 
terms of the quantum numbers. For that purpose the model 
atoms of the older period with their electron orbits have 
still their uses, but the clear cut planetary picture seems at 
the moment to have the same doom confronting it as the 
ether not so much actual denial as mere apathy. Useful, 
perhaps, as mental helps for those without the necessary 
mathematics, but beyond that of little use. 



CHAPTER XVI 

DISTBIBUTION OF A SYSTEM IN ENERGY 

16.1 Quantisation and a priori Probability. When dealing 
with complexions and statistical states of a molecular system, 
we introduced as a fundamental hypothesis the statement 
that all complexions have the same a priori probability which 
involves in its turn the assumption that the representative 
point of a given molecule is as likely to be in one phase -cell 
as another. That of necessity implies that the phase-cells 
have all the same magnitude. Similarly in dealing with 
Planck's oscillator and deriving his expression for the 
average energy in section (14 . 3), we obviously assumed that 
an oscillator is as likely to be on one quantum path as 
another ; these paths take the place of the equal-sized phase- 
cells in the classical treatment of non-quantised systems, 
and again all complexions have the same a priori probability. 
So long as a system does not involve both quantised and non- 
quantised elements, no necessity to fit the two aspects 
together arises ; nor is it even troublesome to deal with a 
system of complex molecules whose positions and trans - 
latory movements follow classical laws and whose internal 
oscillations follow quantum laws. 

But suppose we have to deal with the ejection of electrons 
from atoms in the photo-electric effect or ionisation caused 
by collisions or X-rays ; or dissociation where atoms are at 
times bound as parts of larger particles and at times are free 
and independent particles themselves ; or sublimation of 
molecules from a solid state. In these examples we see the 
possibility of a system being composed of particles any one 
of which may be in a quantised path sometimes and in a 
non-quantised path at other times. In the latter condition, 
division of a phase-diagram into phase-cells is the suitable 
machinery for counting complexions ; in the former not so ; 

151 



152 STATISTICAL MECHANICS FOR STUDENTS 

quantum paths must be used. How are we to combine the 
two ideas ? Even if we regard all quantum paths as equally 
likely between themselves, and all phase-cells as equally 
likely among themselves (i.e., that a particle is as likely to 
be in one phase-cell as in another if it is behaving in a non- 
quantum way, and on one quantum path as another if 
behaving in a quantum way), what is the chance that a 
particle will be in a given phase-cell as against the chance 
that it will be a given quantum path ? Clearly the answer 
will depend on the size of the cell ; the smaller it is the less 
the relative chance. The following procedure has been 
adopted on the grounds of its plausibility, and no facts are 
known with which it is at variance. Recalling as a simple 
illustration the case of Planck's oscillator, we see that a 
given quantum path concentrates on itself as it were, all the 
representative points which in classical conditions would be 
dotted about in the elliptical annulus between it and the 
next path or rather in the adjacent halves of the two neigh- 
bouring annuli which it separates. This area, whose points 
are thus swept into one linear channel, is h. Now if s is the 
area of a phase -cell, and S the area of the whole phase - 
diagram determined by the volume and total energy of the 
whole system of oscillators, then the a priori probability 
that a representative point, if classical motion were involved, 
would be in a given phase-cell, would be s/S. So it seems 
plausible to assume that the a priori probability that a 
representative point shall be on a given quantum path is 
h/S. Thus the relative probabilities for non-quantum and 
quantum possibilities is s/h, or, if we wish to avoid bringing 
in A priori probabilities explicitly, and content ourselves as 
hitherto, by counting complexions and making the number 
stand for the relative probability of a statistical state, we 
must assume that s is equal to h, i.e., choose the phase-cells 
suitable for complexion-counting in non-quantised motions 
to have the size h. 

This is for a system whose particles have one degree of 
freedom and whose phase-diagram is partitioned into areas 
which represent action. For a system whose particles have 
the usual three degrees of motion, it is a natural generalisa- 



DISTRIBUTION OF A SYSTEM IN ENERGY 153 

tion to take as the size of a phase-cell for non-quantised 
motions A 3 , and then to regard the representative point of a 
given particle to have the same A priori probability of being 
in non-quantised motion in a given phase-cell as of being in 
quantised motion on a given quantised path. Alternatively, 
if p is the a priori probability of the point being in six- 
dimensional phase-cell of size s, the a priori probability that 
it is on any quantised path is p h*/s. Will the reader please 
bear in mind that in this we have not been considering the 
particles as having internal degrees of freedom themselves ? 
They have been regarded as simple structureless particles 
capable of flying about in certain parts of the enclosure, 
just like the constituents of a gas, or, on the other hand, 
being " bound " to some centres of attraction, in other 
parts, and so probably subject to quantum conditions. The 
question of complex molecules does not involve so seriously 
the doubt we have referred to. The phase-diagram will then 
have a higher dimensionality than six ; those extra dimen- 
sions which are required to deal with internal motions will, 
as far as we can judge, involve quantum methods through- 
out. There is a difference between the simultaneous exis- 
tence of quantum conditions and non-quantum conditions 
for different degrees of freedom, and the alternation of the 
same degrees of freedom between periods of quantum 
motion and periods of non-quantum motion. 

16 . 2 Energy- Hypersurf aces and Energy-Shells in the Phase 
Diagram. In dealing with the structureless particles of the 
previous section, nothing has been said as regards the 
*" shape " of the phase-cells. In earlier chapters it has been 
understood that the cell is analogous to a rectangle in a two- 
dimensional phase-diagram ; i.e., it is looked upon as an 
extension-in-phase, such that the co-ordinates and momenta 
corresponding to any phase in it are respectively greater than 
some set of values x, y, z, , 7?, , and respectively less than 

* + Sx, , + 8, where Sx, S are small 

increments. But there is no compulsion to adopt this view. 
The fact that in the most probable state the numbers in 
each cell involve e" 1 " as a factor suggests an alternative 
method of delimiting the cells which will prove convenient 



154 STATISTICAL MECHANICS FOR STUDENTS 

at times. The energy of a particle is given by the function 
( 2 + f] z + 2 )/2w, at all events in the absence of an external 
field (and ignoring also for the moment any internal energy 
or energy of rotation). All phases whose energies are 
individually less than are bounded by a region in the phase- 
diagram, such that 

fa + ,a + a = 2m . . . . (16.2.1) 

Its six-dimensional magnitude is determined by inte- 
grating dx dy dz d drj d throughout the volume of the 
system of particles as regards x, y, z, and throughout a 
sphere determined by (16 . 2 . 1) as regards , 17, . The result 
is 

t 7 (2m )*v . . . . (16.2.2) 
*$ 

where v is the volume of the system. The phases which 
satisfy (16.2.1) exactly, have a certain extension-in-phase, 
but it should be realised that it has not the full dimen- 
sionality (six) of the phase-diagram. Just as in ordinary 
geometry, the points whose co-ordinates satisfy an equation 
such as 

f(x, y, z) = c, 

have an extension which is only superficial, and do not 
occupy a three-dimensional volume, so the extension of the 
phases satisfying ( 1 6 . 2 . 1 ) is only five-dimensional. We call 
(16 . 2 . 1) a " hypersurface." * Apart from names, however, 
we realise that on strictly classical lines the chance that the 
phase of any particle in the system satisfies (16.2. 1) is zero. 
But if we consider another hypersurface 

2 +i? 2 + a = 2m( + 8c) . . (16.2.3) 
the phases which satisfy any equation 
P + i* + ? = c, 

where c has any value between 2 m and 2 m (e + Se) 
occupy a six-dimensional region in the phase-diagram, and 

* Of course (16 . 2 . 1) would represent an ordinary surface in a simple 
three-dimensional momentum-diagram. In the phase -diagram, however, 
six dimensions are involved. The absence of x, y, z in ( 16 , 2 . 1 ) shows that 
the hypersurface has " cylindrical " properties. 



DISTRIBUTION OF A SYSTEM. IN ENERGY 155 

the chance of a particle in the system having one of these 
values is not zero ; it is proportional to the size of this 
"energy-shell lying between" the hypersurfaces (16.2. 1) 
and (16 . 2 . 3). The picturesque phrase " lying between " 
has a convenient brevity ; its meaning is quite definite, and 
can be expressed analytically as above. The size of this 
shell can be obtained as the differential of (16.2.2); it is 

27r(2m)He i S(: . . . . (16.2.4) 

Shells such as these lying between adjacent members of 
the family of hypersurfaces (16.2. 1) for varying values of e 
can be selected as phase-cells. In the most probable state 
of the system of structureless molecules the number of 
particles whose phases lie in such an energy-shell is 

Cve-KJS* . . . . (16.2.5) 
where C is determined by 



r 

Jo 



Cv e* e~*" dt = n. 

Jo 

When we pass to the question of internal degrees of free- 
dom, the matter can be treated in an analogous way. 
Leaving aside in its turn and for the moment the question 
of the general position and motion of a molecule, suppose 
the internal structure of the molecule to be defined by 

values of certain co-ordinates q^q^ , ?/ and momenta 

p l9 p 2 , pj. These can at the moment, pending a 

fuller account of the dynamics of such systems, be thought 
of as a suitable number of Cartesian co-ordinates or polar 
co-ordinates of sub-particles with reference to origin and 
axes in the molecule, and the linear momenta or linear and 
angular momenta accompanying them. The salient point 
to bear in mind at present is that the product of any co- 
ordinate and its corresponding momentum has the physical 
dimensions of action (energy X time). The internal energy 
of a particle will be given by some function of the q r and p r , 

<(?i> ?/, Pi, , p f ), or briefly <(#, p). In the 

2 /-dimensional phase-diagram, the phases satisfying 

>) = c (16.2.6) 



156 STATISTICAL MECHANICS FOR STUDENTS 

occupy a 2/ 1 dimensional extension. The 2 /-dimen- 
sional region which " lies between " (16 . 2 . 6) and 

*(<!, #) = * + . (16.2.7) 
contains those phases which satisfy any equation such as 

4(9, P) = c 

where c has a value between e and + Se. It will have a 
magnitude 

X() ^ 

where x(e) is & function of e, involving of course such con- 
stants as are present in the functional form of (f>(q, p). In 
the most probable state of the system of molecules the 
number whose internal co-ordinates and momenta will at 
any moment lie in the energy-shell between (16 . 2 . 6) 
and (16 . 2 . 7) are given by an expression involving 

e~>"x(e)Se. 

16.3 Energy-Hypersurfaees and Paths. Following the 
representative point of any particle, this will travel about 
the phase-diagram in a fortuitous sort of way. For a time it 
may remain on an energy-hypersurface ; it is experiencing 
no influence from other particles or radiation. But at times 
its path will be stretched from some point on one hyper- 
surface to a point on another ; it is then under such influences. 
Let us consider a little more closely the paths when the 
particle is undisturbed and still, for convenience, confine 
ourselves to the internal motion. The co-ordinates q and 
momenta p can theoretically be determined from the 
equations of motion. They are then expressed, each as a 
function of the time t and certain constants.* Of these 
constants a certain number will depend on the internal 
structure of the molecule ; we shall denote them by 

a l9 a 2 , Others, 2/, in number, are introduced in the 

integration of the equations of motion. These integration 

constants we shall denote by b l9 6 2 > &2/- We then 

have 

* Chapter XXIV. will give a fuller explanation of these statements which 
may be accepted now without question. 



DISTRIBUTION OF A SYSTEM IN ENERGY 157 

q l = B l (t, a, b) 
(16.3.1) 

q f = f (/, a, b) 

Pi = 0i (^ , 6) 



jp, = 0, (J, a, 6) 

where flj ( ), fa ( ) are 2/ functional forms. If we 

give a definite set of values to the b constants, then the sets 

of values acquired by q v , PJ, as we vary t from 

oo to +00, constitute a path of the particle in undis- 
turbed motion. The b constants vary from path to path. 
But the a constants have the same values in all paths ; they 
are characteristic of the molecule itself, and its particular 
structure. We represented the energy by <(#, p). If in this 
function we substitute for the q r and p r the functions O r ( ) 
and iff r ( ), the terms involving t must vanish identically, 
since the energy of the particle is constant in undisturbed 
motion, and <j>(q,p) is transformed into a function of the 

a r and 6|! constants, say (a v a 2 , b l9 6 2 , ). 

All this is briefly and picturesquely expressed by saying that 
any path " lies on " some energy -hypersurf ace. If the 
energy is e, then, of course, 

<A K, a 2 , b l9 6 a , ) =- e . (16.3.2) 

Now, in general, this equation can be satisfied by an 
infinity of different values for the set of constants b l9 b 2) 
, b 2 f J for (16 . 3 . 2) is only one equation in 2/ quan- 
tities &x, 6 a > b 2 j. Thus on one energy-hypersurface 

there is in general an infinity of undisturbed paths of the 
particle. Of course it might happen that only one of the b r 
constants appears in the function iff (a, b) y and if it appears 
in a term of the first degree, there is only one path. Never- 
theless, whether the paths are numerous or not, there is no 
guarantee that they " fill up " completely the extension (of 
dimensionality 2 / 1) of the hypersurf ace. Giving every 
value from oo to + oo to t, and all the suitable values to 

&!, 6 2 , b 2 f, we obtain an infinity of sets of values of 

?i, > <?/>#!> > P/> satisfying 

^ (?> P) = c, 



158 STATISTICAL MECHANICS FOR STUDENTS 

but that is no proof that we have thereby obtained all the 
sets of values which satisfy this equation. This point should 
not be overlooked. It has an important bearing on the 
arguments which will be advanced later to justify the 
postulates of statistical mechanics. 

16.4 Quantisation of the Paths. Degeneracy. Still con- 
fining ourselves to the internal motion of the molecule, we 
know that any path* is characterised by the values of the 
/ action-integrals which are derived in the analysis of the 
motion into its latent periodic elements. These are, in fact, 
functions of the constants, and, of course, will vary from 
path to path. Also the energy can be expressed as a function 
of these / action-integrals. 

Quantisation of the paths is effected by selecting those 
values of the action-integrals which are whole-number 
multiples of h. This obviously selects certain energy - 
hypersurfaccs as relevant in quantum atomic physics, the 
others being irrelevant. The journey of the representative 
point from one of the quantum hypersurfaces to another in 
a quantum jump is an occurrence whose details are not 
disclosed by any part of the theory of stationary states. It 
is regarded as occupying a time so brief that only one 
characteristic of it comes into the discussion, the amount of 
energy emitted or absorbed. The contrast to classical theory 
where the journey from one hypersurface to another during 
a period of disturbance is supposed to agree with the laws 
of dynamics, is marked. In quantum considerations, it is 
the quantum hypersurfaces which are all important, and 
the quantum paths on them. Interesting questions will 
arise as to the number of quantum paths on one energy- 

hypersurface. If in E(J l5 J 2 , , J^) we substitute 

r-Ji, rji, ,r f h for the several J r , we obtain a function 

c(r l9 r 2 , TJ) of the quantum-numbers of the path, 

which is equal to the energy corresponding to the hyper- 
surface on which it lies. Now if the r g variables were 

* The internal motion is naturally visualised as the motion of several 
submolecular particles (atoms, electrons) in orbits ; but do not confuse 
" path " with any particular orbit. The path is a synthesis not only^of the 
geometric features of these orbits, but also of the actual conditions of 
movement of the sub -particles in them. 



DISTRIBUTION OF A SYSTEM IN ENERGY 159 

capable of taking any values, then in general many quantum 
paths would lie on a hypersurface, but as they are integers, 
we cannot maintain this statement ; for we cannot assert 
that in general an equation like 



can have a solution in integers at all for a given value of c, 
and if it has one such solution, we cannot therefore assert 
that it has any more such solutions than this one. Still, 
there are certain easily-defined cases in which there are more 
quantum paths than one on any of the quantum hyper- 
surfaces. Thus, if the energy function E(J) happens to be 
a function of 3i + J 2 , J 3 , J 4 ...... , J/, then e(r) is a 

function of r l -\~ r 2 , r 3 , r 4 , ...... r f . Thus any pair of 

integral values of r a and r 2 which have the same sum will 
yield the same energy. But different values of r l and r 2 , 
even if they have the same sum, when combined with one 
set of values for r 3 , r 4 , ...... , r f , correspond to different 

quantum paths which are therefore on the same energy - 
hypersurface. It will be recalled that the fundamental 
frequencies of the internal motions in the molecules are given 

by 

3E(J) 



v, 



so the condition of affairs just mentioned means that v l = i> 2 , 
or two of the fundamental periods are the same. In reality, 
E(J) is a function of only / 1 variables, since the single 
variable, J, + J 2 , takes the place of J l and J 2 . We cannot 
enter into the dynamical theory of this matter here, but the 
reader will probably not find much difficulty in accepting 
the statement that the motion appears to be one in which 
there are not /, but only / 1, degrees of freedom. It is 
said to be " degenerate " on that account ; the question is 
one of importance in statistical theory, as we shall see 
presently. As a matter of fact, any linear relation between 
the frequencies, such as 

a i v i+*2 v 2 + + ajv f = Q . . (16.4.1) 



160 STATISTICAL MECHANICS FOR STUDENTS 

where the a f are positive or negative integers, produces 
degeneracy.* The internal motion in the molecules appears 
to be deprived of one of its internal degrees of freedom ; but 
what is more germane to our purpose is that there are in 
such cases a number of different quantum paths on the same 
energy-hypersurface. If two relations, such as (16 . 4 . 1) hold, 
the system is doubly degenerate ; the molecule appears to 
have lost two of its internal degrees of freedom, and the 
possibility of multiplicity of paths on the same hyper- 
surface is increased. 

We saw that in statistical-mechanical theory the quantum 
paths take the place of equal-sized phase-cells in counting 
complexions. Thus in the most probable state of the system 
of molecules the number of molecules which are in the 
quantum-state denoted by the quantum -numbers r l9 r 2 , 
...... , ry, is proportional to 



If now there are w 8 quantum paths on the energy -hyper- 
surf ace corresponding to an energy e g of the molecule, then 
in the most probable state of the system the number of 
molecules having the energy 8 is proportional to 



If classical conditions held for the internal motion, it was 
pointed out that in the most probable state the number of 
molecules having a definite energy, would be really zero, 
i.e., statistically. A statistical statement could only be 
made concerning molecules whose energies lie between and 
c + Se ; the number is 

- C er^ x(<0 Sc 

The corresponding full statement for the quantum con- 
ditions is that the number of molecules which have an 

* See the Appendix to the Chapter. 



DISTRIBUTION OP A SYSTEM IN ENERGY 161 

assigned energy e consistent with one or more stationary 
states is 

C w er*< h f 

where w is the number of states having this energy. The 
expression wh* replaces x(<0 Se. Both have of course the 
same physical dimensions, viz., those of the / th power of 
action. 



APPENDIX TO CHAPTER XVI 

IT will be less troublesome to follow the discussion on 
degeneracy if we consider the condition (16.4. 1) for three 
degrees of freedom. The reader can easily make the exten- 
sion to any number. In that case there are three action- 
integrals, and the condition 

04 V-L + a 2 v 2 + a s V 3 ~ 

where 04, a 2 , a 3 , are three integers, positive or negative, 
implies that 

8E(J) 8E(J) 8E(J) __ 

a l -X-f + a 2 --7 + a 3 -3-7 V- 

Oj l C J 2 (7 J 3 

Now introduce the substitution 

I 1 ai Jj -f- 12 J 2 ~f~ ^13 JB 

"2 == a 21 ^1 "f" a 22 ^2 ~t" ^23 ^3 

I 3 = a 31 eTj + a 32 J 2 + a 33 J 3 

where the a rs are integers. We can solve for the J r in terms 
of the I r and substitute in E(J) ; the result is a function of 
Ij_, I 2 , I 3 , say F(I). It is easy to see that 

8E(J) = 8F(I) 3Ii 8F(I) 8^ 2 8F(I) 8I_ 3 

8J X " 8Ii *8Ji 8I a "8Ji 8I 3 '8Ji 

8F(I) 8F(I) 



and similarly 

3E(J) __ 8F(I) 8F(I) 8F(I) 

8E(J) 8F(I) , 8F(I) , 8F(I) 

-^r~- == a !3 -3T~" + ^23 -~f- + a 33 ~-^f 

dJ 3 oL l ol 2 o! 3 

8.M. 



162 STATISTICAL MECHANICS FOR STUDENTS 

So that the condition for degeneracy, is satisfied if 

011 <*1 + 012 a 2 + 013 a 3 = 

a 2l a 1 + 022 a 2 + 023 a 3 
a 31 04 + a 32 a 2 + a 33 a 3 = 0. 

Now if we choose any two integers m and n, we can find 
another number p such that 



If p is not an integer, it is nevertheless a commensurable 
number since a v a 2 , a 3 are integers ; hence, by clearing of 
fractions we find three integers b ly & 2 , ^3> suc h ^hat 
b 1 a 1 + 6 2 a 2 + 6 3 a 3 = 0, 

and from the method of finding them it would appear that 
there exists an unlimited number of ways of doing so. Thus 
the three conditions for degeneracy written above are 
perfectly feasible, but they would appear to be only three 
of an unlimited number of such conditions. As a matter 
of fact there is really one too many. It is a well-known 
theorem in simultaneous equations that we cannot satisfy 
three equations 

a n x + a l2 y + a 13 z = 

021 X + a 22 V + 23 * = 
31 X + a 32 y + 33 Z = 

simultaneously for any arbitrary values of the nine co- 
efficients. If they are to be true simultaneously, then it 
must also be true that 

3i = A a n + /A a 2l 

32 = A 012 + p 022 

a 33 = A a 13 + V* 023 
where A and /z are some two multipliers. Thus it appears that 

J s = A Ii + p I 2 , 

and thus we have reduced E(J X , J 2 , J 3 ) to a function of 
two variables, I and I 2 , since I 3 is a linear function of these 
two. The mechanical system then behaves as if it had only 
two independent action-integrals, two independent periods, 
and only two degrees of freedom. 
When it comes to quantisation, we see that the energy, 



DISTRIBUTION OF A SYSTEM IN ENERGY 163 

being expressible as a function of I x and I 2 , is expressible 
also as a function i/t(p, a) of two integers p and a, which are 
given by 

P = 11 r i + 12 **2 + 013 7*3 

a == a 21 r x + a 22 r 2 + a 23 r 3 

where r l9 r 2 , r 3 , are the quantum numbers of a path. Clearly 
this is compatible with the same energy for different sets of 
values for r l9 r 2 , r 3 . Thus the quantum numbers r^ + a l9 
r z + a 2> r s + a s> would give the same p and or, or for that 
matter, r^ + ka l9 r 2 + fca 2 , r 3 + ^3, would do so likewise 
where k is any integer. Still the choice is not unlimited ; 
for we must bear in mind that the quantum numbers must 
be positive integers, while a l9 a 2 , a 3 are not all positive. 
This condition will, of course, limit the choice of suitable 
values of k. 

It may be of interest to point out the general sort of 
evidence which experiment can give as to existence' of 
degeneracy in atomic systems. The most notable example 
is the Zeemann effect. A spectral line of an atom is split 
by a magnetic field into a number of lines with slightly 
different wavelengths. This means that the mechanism in 
the atom has several different elements of periodicity which 
have the same frequency. The magnetic field affects each 
element rather differently, and the alteration of frequency 
produced is not quite the same in each case, and so each 
element gives unambiguous evidence of its presence. 



CHAPTER XVII 

QUANTUM THEORY OF THE SPECIFIC HEATS OF GASES 

17.1 Specific Heat and the Internal Oscillations of the 
Molecules. From the considerations outlined in the previous 
three chapters, we find that the internal motions of the 
molecules contribute to the energy-content of the system 

C 27 , w 8 e-^* 

where the summation is over all the energy hypersurfaces, 
and the constant C is determined by the equation 

C Sw s e~^* = n. 

Let us represent the lowest quantum energy state by 
the suffix zero, and put C e~^ = C . By Bohr's second 
postulate, if a quantum jump takes place from the hyper- 
surface 8 to the lowest quantum state, the frequency of the 
radiation emitted is i> 8 where 

, Co= hv 8 . 
Thus 

C (1 + 27 w 8 e-^"') = n, 

8 = 1 

and the energy-content is 

C, (e. + 27 c. w 8 e- hv <). 

=i 

Combining these, we see that the energy-content is 
n . e + 2 c. w s e- h "> 

I +Zw t *->*> ' ' ' I"- 1 ' 1 * 

and if we differentiate this expression with respect to 0, we 
obtain the contribution made by the internal motions of the 
molecule to the specific heat of the system of molecules. If 
we do so, we obtain an expression whose numerator is 

nh(Zv 8 w 8 e-* hv (1 + 27 w e-^'w 



164 



SPECIFIC HEATS OF GASES 165 

and whose denominator is 

(1 + 2 w, e-'*"') 2 . 

The latter is positive and greater than unity, so the con- 
tribution to the specific heat turns out to be less than 
nhjkd 2 multiplied by the sum of a number of terms such as 

(v e A- v v e v } w w e~^ v 8+ v u) 

\ V 8 e # I V U *U V U t S V 8 e / w W U ^ 

This latter expression is equal to 



Thus the contribution is less than 

/ __ \ 2 

vi z* y y i s u i ID in 0~t jL ' i ( i '8^~ i 'ui 
nk ^^(~j: r ) w > w ^ 

Now the frequencies involved are according to spectro- 
scopic evidence in the ultra-violet, luminous or high infra- 
red ranges of the spectrum ; their order of magnitude is 
certainly not lower than 10 14 . That being so, it will be easily 
seen that for ordinary temperatures the value of the index 
in any exponential factor is numerically greater than 15, 
and in the majority very much greater. But e~ 15 and other 
experimental factors still smaller, will practically render the 
multiplier of nk in the previous expression negligible. The 
smallness of the contribution is still more marked for low 
temperatures. Only for very high temperatures would the 
internal degrees of freedom begin to contribute an appre- 
ciable fraction of nk or R for each degree of freedom in the 
internal structure of the molecule. In this way the quantum 
theory surmounts one of the difficulties of the older theory, 
which required for each degree of freedom a contribution to 
the specific heat of R/2 as regards kinetic energy and an 
amount of the same order of magnitude as regards potential 
energy. In general terms, and putting aside the rather 
clumsy expressions involved in the analysis, the spectro- 
scopic evidence shows that the great majority of the mole- 
cules will be in the lowest internal quantum state, by reason 
of the smallness of the factors, such as e~^ s . Thus the 



166 STATISTICAL MECHANICS FOR STUDENTS 

internal energy will not differ much from n . At a some- 
what higher temperature there will be a very small change 
in the distribution of the molecules among the higher 
quantum states ; the exponential terms are not seriously 
affected ; practically the internal energy will still be H O , and 
the quotient of the change by the rise of temperature is 
insignificant. Only at extremely high temperatures when kQ 
would be approaching in value to some of the hv 8 , would the 
change in internal energy with rising temperature begin to 
show itself in the specific heat. 

17.2 Molecular Rotation and the Specific Heat of a 
Diatomic Gas. The order of magnitude of the frequencies 
involved in electron motions within the atoms or even in the 
relative vibrations of the atoms constituting a molecule is 
therefore so great that no contribution to the specific heat 
of a gas can be expected from this quarter except at extremely 
high temperatures. Yet we know that the ordinary trans- 
latory movements of the gas molecules can only contribute 
an amount 3R/2 to the thermal capacity of a body of gas, 
and this is too small under ordinary circumstances except 
for monatomic gases. We saw in Chapter V. that in 
the case of diatomic gases the balance, approximately R, 
could be accounted for reasonably enough by assuming a 
further two degrees of freedom involved in a rotation of 
the m'olecule about an axis at right angles to the line joining 
the centres of the two atoms ; rotation about this latter line 
itself is excluded from consideration for reasons stated in 
that chapter. Nevertheless experiments, especially on 
hydrogen, show that this contribution does not remain even 
approximately near R as the temperature decreases ; the 
evidence is very much in favour of the view that as the 
temperature approaches absolute zero, the contribution of 
molecular rotation to the specific heat diminishes asymp- 
totically to zero. This suggests at once that in some way 
the rotational motion is quantised and not subject to classical 
equipartition, while the frequencies involved are of such 
magnitude that the average energy must be close to k6 for 
ordinary temperatures, but much less as 6 decreases below 
normal values. Now the average energy is 



SPECIFIC HEATS OF GASES 167 

hv 

exp (hv/kO) 1 

or 7 /j x 

e*~^T 

where x = hvjkO. Putting 9 = 300, k = 1-37 X 10"" 18 and 
h = 6-55 X 10~ 27 , we see that to make x = 1, v must be of 
the order 6 X 10 12 . Consequently frequencies between 10 12 
and 10 13 would, at ordinary temperatures, approximately 
involve the classical partition of energy since for such condi- 
tions xj(e x 1) would approximate to unity. But at much 
lower temperatures x would increase to the order of magnitude 
of 10 or even 100, and then x/(e x 1) is practically negligible. 
Now, as a matter of fact, it has been discovered that in the 
absorption spectrum of water-vapour there actually exist 
some lines whose frequency are in the region about 5 X 10 12 . 
These lines are usually referred to as the rotation-spectrum ; 
and as a piece of additional evidence there has also been 
observed a well-marked effect produced by rotations of such 
frequencies on the lines in the infra-red of order 10 14 , which 
arise from the atomic vibrations in the molecules of water- 
vapour and the hydrogen halides. 

To be sure we are not dealing here with an oscillator in the 
Planck sense, having one frequency for all amplitudes, and 
the treatment just outlined is too inadequate. But a 
method of quantising a rotation, easily deduced from 
the considerations of Chapter XV., forms the basis of a 
theory which accounts for the behaviour of diatomic gases 
as regards their specific heats, apart from some minor 
discrepancies, which appear to arise rather from our lack of 
knowledge of the actual structure of the molecules than from 
any serious deficiency in the theory. 

A diatomic molecule pictured as a dumb-bell has the line 
joining the atoms as one principal axis of inertia. Any two 
lines at right angles to this and to each other and passing 
through the centre of gravity can be taken as two other 
principal axes of rotation. The atom is a " symmetrical 
top/' but will the reader carefully bear in mind that rotation 
about the axis of symmetry, i.e., the line joining the atoms, 



168 STATISTICAL MECHANICS FOR STUDENTS 

does not come into consideration ? One reason for excluding 
this rotation was given earlier in the classical treatment ; 
another reason, more in keeping with quantum ideas, will 
emerge presently. It is with rotations about axes at right 
angles to the line of symmetry that we are concerned. Let 
A be the symbol for the moment of inertia of the atom about 
such an axis. This is as a matter of fact equal to M l r x 2 + 
M 2 r 2 2 where M x and M 2 are the masses of the atoms, and 
r l and r 2 are their distances from the centre of gravity, and 
it is not difficult to show that this is also equal to M a 2 where 
a is the distance separating the atom-centres and 

JL--L + 2-. 

M M 1 M 2 

The angular momentum round the axis is Aw if co is the 
angular velocity. The action accumulated in one period, 
i.e., in one rotation is then 2 TT A o>, and so, according to 
quantum views, rotations with any angular velocity cannot 
exist for an interval of undisturbed rotation, but only 
rotations for which the action-integral 2 n A co has a value 
such as hy 2 h, 3 h, etc.* Hence the " rotator " can only 
have one of a series of discrete angular velocities co 1? o> 2 , 
cu, ...... where 



When rotating with one of these velocities it is in a quan- 
tum state defined for the moment by one quantum number r. 
But one number is really insufficient. We have two degrees 
of freedom to deal with (the third, rotation around the axis 
of symmetry, has been definitely excluded), and we must 
have two quantum numbers. We have here, in fact, an 
example of degeneracy, and in the early days of the theory 
this feature of the situation caused some trouble for a reason 
which will appear presently. Leaving this ambiguity on one 
side for a moment, we see that the energies of the quantum 

* The reader will observe that angular momentum being the product 
of mass, distance squared and angular velocity actually has the same 
physical dimensions as action. 



SPECIFIC HEATS OF GASES 169 

states of the rotator defined by (17. 2. 1) are given by 
equations such as 



m <17 - 2 ' 2) 

Thus the energy in a quantum state varies as the square of 
the quantum number, a rule quite distinct from the rule 
which holds for a Planck oscillator ; in that case the variation 
is with the first power. We would now be in a position to 
obtain the average energy of a molecular rotator in the 
system if we knew the number of quantum states consistent 
with this energy, for as we have seen, it is equal to 



2 w 



r 



2 w 



(17.2.3) 



One of the earliest attempts to use this expression was 
based on the assumption that the w r are each unity and that 
the result thus obtained should be doubled so as to take 
account of the two degrees of freedom really involved in 
rotation about an axis at right angles to the symmetry-line 
of the molecule. It will be instructive to carry out this 
evaluation before proceeding to later attempts to cope with 
the difficulty. On such a view the rotational energy of the 
system is 

00 

2 e e'^r 



2w l 

d fJL 

fj ( 

2n log 1 



, 
where 



a = 



170 STATISTICAL MECHANICS FOR STUDENTS 

The rotational part of the thermal capacity of the n mole- 
cules of gas is then equal to 



dO dfju 

5 l^rr F /,-of* 



At low temperatures //,, and therefore a, are relatively large. 
The series S e~ ar * practically reduces to 1 + e~ a , and the 
logarithm of this to e~ a ; thus the rotational thermal 
capacity is equal to 

,72 

- 



and since for large values of /x or a, the factor e~~ a " swamps " 
ju, 2 , this has a limit zero as /z increases indefinitely. For high 
temperatures a is small and since 

1 + e~ a + e~ 4a + e- 9a + ...... 

= a -l{ Xl + e -V (x 2 -xj+e -*>\xz - x 2 ) + e~*>* 



where x 1 a*, x 2 = 2a*, x 3 3a J , etc., the series S e~ ar * 
can be approximately written as a definite integral 



f 00 
a~M e~ x * dx, 

Jo 



which is ^ (?r/a)*. Thus at high temperatures the rotational 
thermal capacity has as its limit 



which is equal to 



The conclusions concerning the extreme limits are quite 
satisfactory ; but, unfortunately, a closer investigation shows 



SPECIFIC HEATS OP GASES 171 

that the thermal capacity rises to a maximum above R and 
sinks to a minimum below it before reaching the limit R. 
The experimental observations do not bear out this con- 
clusion. Various other suggestions were made for dealing 
with the doubtful situation arising from the degeneracy of 
the motion, but all suffer more or less from the defect 
mentioned above, viz., a tendency for the computed value 
to rise somewhat above the experimental value. Still they 
are an improvement on the earliest result obtained by the 
mere doubling method. The way in which the degeneracy 
of the motion is removed involves a rather wider knowledge 
of dynamical science than is assumed in this book. It must 
suffice to say that one plausible suggestion leads to the 
conclusion that there are really r different quantum states 
with the energy r 2 k 2 /S n 2 A , and so w r should be put equal 
to r. Incidentally, this implies that in no state is the energy 
of rotation absolutely zero, the lowest quantum state has 
the energy of rotation h 2 /S7T 2 A. This feature, viz., that 
the lowest quantum state is not one devoid of rotational 
energy, is a common feature of all the hypotheses used to 
explain not only the thermal behaviour but also the results 
of spectroscopic analysis. In the same manner as we pro- 
ceeded above, we now find that the rotational thermal 
capacity of n molecules is 

72 CO 

R jLt 2 - log 27 (r e~ ar *) 

At low temperatures this vanishes in the limit and at high 
temperatures the series approaches the value of the definite 

integral 

/<# 

x * dx, 



_ f 
a" 1 ! x 

Jo 



i.e., l/2a, which gives us a value R for the above expres- 
sion at the upper limit of 0. But computation of the 
series still shows a discrepancy at moderate values. Thus 
about 200A the computed value for hydrogen is -82 R 
as against the experimental -72 R, a rather serious differ- 
ence in view of the claims made by the experimentalists 
as regards the precision of their measurements. By the 



172 STATISTICAL MECHANICS FOR STUDENTS 

time, however, we reach the temperatures of our normal 
surroundings, the discrepancy disappears, the values at 
freezing point of water being -936 R and '937 R respectively. 
The reader may recall the statement made earlier that 
recent spectroscopic work calls for " half -quantum numbers " 
in certain cases, meaning that although the quantum 
numbers increase by integral amounts from quantum state 
to quantum state, they are not necessarily integers them- 
selves, but may have values such as '5, 1-5, 2-5, etc. This 
suggestion has been tried in this problem, and r being treated 
this way, it is found that w r can either have the series of 

values 2, 4, 6, 8, or 1 , 3, 5, 7, In either case 

this method gives better agreement with experimental facts 
than those based on integral values of the quantum numbers. 
As mentioned above, the source of the discrepancies may lie 
in the assumption of a constant A throughout ; at higher 
temperatures there may be a " stretching " of the molecules 
owing to increased rotation and some change in the moment 
of inertia. The computations incidentally enable us to find 
a numerical value for A ; the values differ somewhat 
according to the method used, but they all agree in giving 
something of the order of magnitude 10~ 41 in C.G.S. units. 
This result is of the same order of magnitude as the values 
obtained from the observation of the rotation spectra and 
rotation-vibration spectra of vapours of water and the 
hydrogen halides, and is consistent with the value 10~ 24 
gram for the mass of the hydrogen atom and -5 X 10~ 8 
cm. for the diameter of the hydrogen atom. The latter 
are derived from the kinetic theory of gases or from Bohr's 
theory of the Balmer series, and it is assumed that the atoms 
in the molecule are close together. 

We can now give the very simple explanation on quantum 
lines why we leave out of consideration rotation round the 
axis of symmetry of the molecule. The moment of inertia 
of the molecule round this axis is much smaller than A . All 
our present knowledge of atomic structure points to the 
conclusion that the mass of an atom is practically con- 
centrated in the small nucleus, whose linear dimensions are 
of a much smaller order of magnitude than the radius of the 



SPECIFIC HEATS OF GASES 173 

atom itself, i.e., the radius of the outer electron orbits in the 
normal state of the atom. This latter radius is the order of 
magnitude of the distances of the atom centres from the 
centre of gravity which are used in calculating A ; but in 
calculating the moment of inertia around the axis of sym- 
metry we would use distances of an order of magnitude 
given by the size of the nucleus. Call this latter moment of 
inertia C. If rotation about this axis came into play, the 
smallest possible angular velocity would be given by h/2irC 
and the corresponding energy by h 2 /$7r 2 C. In comparison 
with h 2 l87T 2 A, this would be very large ; the chance, there- 
fore of a molecule being in the lowest quantum state of 
rotation around the axis of symmetry would be very small 
relative to the chance of it being in the lowest, or even in 
many a higher quantum state of rotation around axes at 
right angles to the symmetrical line. Molecules rotating 
round the axis of symmetry are in consequence too few in 
number to affect the final result. 



CHAPTER XVIII 

THE ELASTIC SPECTRUM OF A LINEAR LATTICE 
OF COHERING PARTICLES 

18 . 1 The "Co-ordination " of a Chain of Particles. For 
a real understanding of the manner in which the quantum 
hypothesis can be applied in the treatment of the specific 
heats of solid bodies, it is necessary to know something of the 
mathematical method by means of which the analysis of the 
irregular heat motions of the atoms into component simple 
vibrations is effected. The reader is probably aware that 
the physical and chemical facts concerning crystalline 
materials support the view that a " molecule " is apt to lose 
its identity in a solid. The constituent atoms of the mole- 
cules are arranged in space lattices, corresponding atoms of 
each molecule forming one lattice, other corresponding 
atoms forming another lattice, the lattices interpenetrating 
one another. Under such an arrangement it is hardly 
possible to say that one particular atom is associated 
with another particular atom to form a molecule. The 
partnerships set up when the substance is liquefied or 
vaporised hardly exist in the solid state. We shall, therefore, 
take as an ideal simple solid a group of particles of one kind 
arranged in a cubical lattice. This will form a convenient 
model for a monatomic solid. 

In presenting the mathematical method referred to, it 
will be advisable to show its application in the first instance 
to something even simpler than the ideal solid, viz., a row 
of particles cohering together by reason of strong attractions 
to form a chain, which, however, is not really rigid, but is 
capable of oscillating and displaying an enormous number of 
forms following one another in a manner determined by 
dynamical laws. 

Every one with a musical ear can detect in the sound of a 

174 



ELASTIC SPECTRUM OF A LINEAR LATTICE 175 

note played on any instrument a series of simple notes with 
definitely related pitches, the so-called fundamental tone 
and its overtones. It is well known that these sound sensa- 
tions are related to vibrations in the sounding body which 
can be analysed into more elementary vibrations each 
having a definite frequency. One of the most customary 
ways of illustrating this fact in works on Acoustics is to take 
the violin or piano string as an example. To apply the 
mathematical method, in its most simple form, we idealise 
these and conceive a string without any physical property, 
except length, mass and the capacity to support a tension 
and a slight stretching without breaking. It is in the first 
instance regarded as a linear continuum, so that the smallest 
fraction of the distance between its ends contains its pro- 
portional amount of mass. When such a string vibrates, we 
can imagine that its form at any instant can be rendered 
permanent for a while, so that we can study it at leisure. 
A cinematograph film of its behaviour, for example, can be 
stopped with a particular picture showing. The possible 
shapes have an infinite variety and complexity, yet they 
can be dealt with in a very powerful and elegant manner by 
a famous theorem due to Fourier. The geometry looks 
absolutely unmanageable, but the analysis is not at all 
difficult to understand. Let us assume that the string is 
fastened at its two ends. Distance from one of these ends 
along the string direction we shall denote by the symbol #, 
the whole length being /. Let us also for convenience assume 
that the vibrations are confined to one plane, i.e., that each 
element of the string has only one degree of freedom. The 
displacement of an element in this plane from its equilibrium 
position, and, of course, at right angles to the direction of the 
string, we shall denote by . The fact that the string has a 
definite shape at all at our moment of observation is expressed 
by saying that is a function of x. In fact, when we write 



we mean this. We have written down the " equation of the 
curve." Yet if we should glance at any of the shapes of 
vibrating strings that have been actually obtained with 



176 STATISTICAL MECHANICS FOR STUDENTS 

violins, etc., we might well be dubious about the possibility 
of finding the actual " form " of the function *jj (x). Yet 
there is a way of doing it which enables us to obtain i/j (x) 
approximately with the possibility of carrying the approxi- 
mation to any degree of accuracy necessary. Formal proof 
cannot be given here, but reference to any standard text on 
the calculus will substantiate the following result 



. . x .770;. . ZTTX , , . TTTX 
yt (x) = q l sin + q 2 sin + + q r sin 

V V L 

+ adinf (18.1.1) 

where the coefficients are definite integrals defined thus : 

TTTX , 



2 ( l . . . . 
ft = y <A ( x ) sm 

I JO 



If the shape were actually drawn on a sheet of paper, we 
could, by multiplying each ordinate by the value of sin 
r TT x/l at the point, obtain a new curve whose area would 
be the coefficient q r multiplied by 1/2. Work of a kind 
analogous to this is actually carried out to-day for 
practical ends, e.g., in analysis of the tides. The reader will 
therefore realise that the statement that we could find the 
equation of the curve to any desired accuracy is not merely 
" theoretical; " it is quite practicable, if somewhat tedious 
at times. The salvation of the practicability lies in the fact 
that in a great many cases the series is so highly convergent 
that a half-dozen terms or even fewer serve very well. (In 
the most highly developed " harmonic analysis " of the 
tides at the present day, investigators seldom use more than 
two dozen harmonic constituents. ) Even such an apparently 
intractable shape as a " zig-zag " is extremely well repre- 
sented by six or seven terms. 

So far this is a matter of geometry and analysis. Now we 
take up the kinematic side of the matter. The string does 
not stay in this shape. Shape after shape follow in a con- 
tinuous succession depending on the tightness of the string 
and the shape from which we " let go." In the analysis this 
succession of shapes corresponds to a succession of values for 
the group of coefficients, q r . Each of these is in fact a function 



ELASTIC SPECTRUM OF A LINEAR LATTICE 177 

of time. If we knew what these functions were, we would 
have in that knowledge summarised the whole history of the 
string's behaviour. In short, #!, # 2 > ...... are " co-ordinates 

of the string." Here we take a more general attitude towards 
the word, co-ordinate, than hitherto. It is not to be merely 
restricted to Cartesian or polar methods of specifying the con- 
figuration of a system. This will, in fact, be a useful illustration 
to fall back on when we come to treat the most general way 
of stating the laws of dynamics in a later chapter. Any set 
of quantities which when known specify precisely the com- 
plete configuration of a system can be regarded as co- 
ordinates of the system. The very name " arranged to- 
gether," suggests this. 

If now we wish to find the functional forms which show 
how the various q r depend on t, we must naturally use the 
laws of dynamics. It is known from the application of these 
that the displacement is a function of x and t which 
satisfies the partial differential equation 



where T is the tension of the string, and M is the mass of 
unit length of it. From this and equation (18 . 1 . 1) it follows 
that 

7r 2 T * . TTTX d\ . TTTX 

-- 27 r 2 q r sin - M sin - 

I 2 r -i I r-i dt 2 I 

If this is to be true for any value of x, it follows that each 
q r satisfies an equation of the type 

d\ 2 7T 2 T 

_ L 7*2 ___ g 

dt* I 2 M ^ 

Hence 

$ = a r sin (w r t e r ) . . (18.1.2) 
where 



and a r and e r are arbitrary constants. The constants o> r are, 
of course, independent of the initial conditions ; they are 



178 STATISTICAL MECHANICS FOR STUDENTS 

determined entirely by the nature of the system. The o> f 
thus form a harmonic series of pulsations (pulsation is 2 TT X 
frequency) whose fundamental is (rrjl) . (TjM)^. On the 
other hand, the a r and the r are the usual integration con- 
stants entering into the solution in the integration of the 
equations of motion. They are arbitrary in the sense that 
if we start the string to vibrate by releasing it from one 
configuration, and then give it another vibration with a 
different initial form, the two sets of a f and r will be 
different. A glance at (18 . 1 . 2) will justify the use of the 
word " amplitude " for the a n but again in a more general 
sense than hitherto, not being confined to the amount of 
excursion from side to side which any element of the string 
makes. To make this clear, let us suppose that we choose all 
the a r to be zero except a l ; all the q r are also zero except q l9 
and we find for the displacement at time t of an element of 
the string situated at x the result 

. . . . TTX 

a l sm (a) l t ej sm . 

i 

If x = or I, this is zero, as it must be since the ends are 
fixed ; if x = 1/2, the vibration is a l sin (coj t j), and so 
a l is the amplitude, in the ordinary sense, of the vibration of 
the middle element of the string ; but for any other position 
this is not so ; the amplitude, in tlie customary sense, is 
a l sin (TT x/l), which gradually decreases to zero as we 
approach either end. The string is then vibrating as a whole 
between two extreme sine forms 

y . TTO; 

=a 1 sm 

, y . TTX 

and = a 1 sm . 

L 

This is its fundamental vibration. If, however, we make 
all the a r zero except a 2 , the displacement is given by 

v i \ %1TX 

= a 2 sm (co 2 t e 2 ) sm . 

i 

This vanishes always at the middle point as well as at the 
ends ; there is a " node " there. Only at the points of 



ELASTIC SPECTRUM OF A LINEAR LATTICE 179 

quadrisection will the amplitude of displacement be a 2 > 
there are two " antinodes " situated at those points. The 
string vibrates as it were in two halves between the extreme 
form 

j. . %7TX 

4 = a 2 sin 

I 

and the extreme form 

2-7TX 



= a 2 sin 



I 



which are two complete sine-curves. The extension of this 
is obvious. The a r are the amplitudes of the q r co-ordinates. 
In the most complicated vibration of the string there are 
latent all the simple vibrations with an ascending series of 
harmonically related frequencies. 

So far we have been dealing with the ideal string familiar 
in works on Acoustics. We must now point out how the 
result is modified if we replace the string by a chain of 
equally -spaced particles. For one thing we cannot be 
involved in an infinite series. If there are / particles, each 
with one degree of freedom, we only require / co-ordinates, 
no matter how chosen, to specify any configuration. So that 
gives us a broad hint that instead of an infinite series we must 
have something like this for an instantaneous form of the 
chain 

u, . . 7T X k . . . 2 7T X,. . , / flTXlf 

* = <i sin * + < 2 sin j- + + ^ sin J ^ 

where x k is the distance of the k ih particle from one end and 
j. its displacement. There is a distance l/(f + 1 ) between 
each particle, assuming that the two end ones are anchored 
to two immovable particles which we can suppose to be 
labelled and / + 1 respectively. Thus the previous 
equation can be written 

v , krr . . . 2Ic7T . . , . fkrr 

t = #i sm jq7j +^2 sm j^-j + + ^/ Sin jq7j 

(18.1.4) 

(Incidentally this makes and g f+ l zero as it should.) 
Now investigation shows that this surmise is quite correct, 

N2 



180 STATISTICAL MECHANICS FOR STUDENTS 

and dynamical analysis * leads to a functional dependence 
of the/ co-ordinates, <f> v ^ 2 , ...... <f> f on t given by 

</> r = a r sin (K T t r ), 

where the a r and r are 2/ arbitrary constants of integration, 
while the K r are / pulsations given by equations of the type 



In this, T is still the tension, i.e., the attraction between two 
neighbouring particles, m is the mass of one particle, i.e., 
M l/f, and I' is the distance between particles or l/(f +1). 

The analogy between the <f> r and the q r or between the 
to r and K r is obvious, and the question arises how far We can 
make use of the mathematical methods, suitable for dealing 
with a continuous medium, for the treatment of a medium 
with a discrete structure not only in the case actually under 
consideration, but also when three-dimensional bodies are 
in question. In the first place 



I'm Im 

_/(/+!) T 
_ .__. 

Further, if r is small compared with /, and / is a very great 
number 

r 77 r TT 

OlT-k ^^3 



and _ r ? 

KT ~~~T \M) 

Thus, provided r is a reasonably small fraction of /, say 
0-1 at most, there is practical equality between K T and to r . 
But as we ascend into the higher ranges of the K T frequencies, 
we cannot take them to be the ascending integral multiples 
of the fundamental frequency. However, we can anticipate 
that an application of the quantum theory will, in view 
of the general feature, which has already been illustrated in 

* See Routh's Rigid Dynamics, Vol. II., Chapter IX., or Rayleigh's 
Theory of Sound, Vol. I., Chapter VI. 



ELASTIC SPECTRUM OF A LINEAR LATTICE 181 

other connections concerning the decreasing significance of 
increasing frequency, turn out to produce results not far 
from the truth if we overlook the departure of the higher 
K T from the simple iile and assume that we are concerned 
with / harmonic frequencies which are all integral multiples 
of the frequency v given by 

1 

v = - 



21 \M 

That being so. we can plausibly assume that the <f> r agree 
with the q r . The assumption is fully justified by a closer 
analysis in the cases where the Fourier scries is convergent, 
and it is only with such cases that we are physically con- 
cerned. 

Thus we have analysed the motion of any of the particles 
in an undisturbed vibration of the chain into simple har- 
monic constituents. If as before x stands for the distance 
of a particle from one end 

~ i . r TT x . IT 77 c t \ 
= S a r sm sin ^ e r j 

where /T\*. 



r V 

\M) 



It will be observed that the form of the chain exactly 
repeats itself when t increases by 2 l/c. This is the period of 
vibration ; it is also the time in which anything travelling 
with a speed c would cover a distance equal to twice the 
length of the string. Indeed it is easily inferred from the 
general theory of the string that c or (T/M)* is the speed with 
which a transverse pulse would travel along an unlimited 
string with this tension and mass per unit length. 

18.2 The Energy of the Chain. The velocity of the 
particle whose distance from one end is x, is given by 

V (i n m % 
= Z q r sin 



. r 77 x . , . . 
= aj r a r sm sin (a) r t e r ). 



182 STATISTICAL MECHANICS FOR STUDENTS 

The kinetic energy is equal to - M I 2 dx, and since 



1 



. r 77 X . S 7T X j - ., . 

sin sin dx = if r =f= 

o I I 

= _ if r 
2 



it is easy to see that the kinetic energy is 

..... (18.2.1) 



As regards the potential energy in any configuration, it is 
the work required to produce the necessary stretching against 
the tension T. If in this configuration an element dx of the 
straight string is stretched to a length ds of an element of the 
curve, the potential energy in that element is T (ds dx) 
or T (ds/dx 1) dx. But 



dx [ \dx 

= f i /dty _i /<2\ 4 , 1 

{ "*" 2 \dx) 8 \dx) J 

Neglecting terms beyond the second power, since d^/dx is 
small compared to unity, the potential energy of the string 
turns out to be 

1 T C l /d\*dx, 

and since 

^ _ ^ P r 7T a: 

dx /r^i f Z 

we can demonstrate as above that this energy is 

7T 2 T * 

2r*q* (18.2.2) 

The kinetic energy and potential energy are by this 
analysis each separated into / parts, any one part being 
associated with a co-ordinate or its " velocity," where the 
word " velocity " is now used to indicate the rate of change of 
a generalised co-ordinate with respect to time and is not 



ELASTIC SPECTRUM OF A LINEAR LATTICE 183 

necessarily the actual rate of movement of any particle. 
The similarity of the mathematical results to those for a 
system of simple oscillators is obvious, and one important 
feature is common to both, viz., the equality of the average 
kinetic and potential energy of the linear lattice not only 
in toto, but also in its analysed parts. The average kinetic 
energy associated with the co-ordinate q l r is by (18 . 2 . 1) 

Ml 2 2 
"8 ^ "" 

The average potential energy associated with q r is by 
(18.2.2) 

V * T 2 2 

-*r r a - 

and since o> r 2 = r 2 (n/l) 2 (T/M), the equality follows. Thus 
from the point of view of mathematical procedure, the 
atomic chain can be regarded as a complex molecule with 
/ internal degrees of freedom, represented by co-ordinates 
q r and velocities q r , each following the simple harmonic law ; 
but it is necessary to be on guard against thinking that these 
are each related respectively to one link in the chain. All 
the q r and the q r enter into the displacement and velocity of 
any individual atom of the chain. 

18.3 The Statistics of a System of Atomic Chains. Having 
thus dealt with the mechanical side of this problem as a 
preliminary illustration for the solid lattice, let us turn to 
the statistical aspect in a similar anticipatory vein. 

Conceive that in a cubical enclosure an enormous number 
n of such chains each containing / atoms are stretched across 
from wall to wall, each anchored to the walls at its ends. 
The enclosure also contains gas molecules, so that the whole 
may be regarded as a mixture of gas molecules and " chain- 
molecules." Exchange of energy goes on between the 
various members of the system, and we can work out in the 
usual manner the conditions of statistical equilibrium. Just 
as we have generalised the meaning of co-ordinate and 
velocity, so we can give a wider significance to momentum ; 
the full import of the step will be more apparent in Chapter 
XXIV, but we define the r th " generalised component of 



184 STATISTICAL MECHANICS FOR STUDENTS 

momentum " to be equal to the partial differential co- 
efficient of the kinetic energy with respect to q r and denote 
it by p r , so that in the present case 

Ml . 

Pr ~ 3r 

We can now represent the 2f quantities q r , p r in a 2/-dimen- 
sional phase-diagram or in/ two-dimensional phase-diagrams, 
partitioning into phase-cells and counting the complexions 
in any statistical state just as before. If the reader feels any 
qualms about this procedure, feeling in a vague way that it 
is hardly as justifiable a process as when we represented 
actual co-ordinates and velocities or momenta of particles 
in the usual sense in a phase-diagram, he can banish them 
without any fear. He will see later that in so far as we can 
accept the procedure to be a justifiable one in the latter case, 
it is equally justifiable in the former. The particular 
character of the laws of dynamics, when given their most 
general mathematical form, takes care of that. Since the 
energy involves only squares of co-ordinates and momenta, 
we will arrive at the conclusion of equipartition of energy 
on the average among the various co-ordinates and momenta 
of the system, an amount | kd to each, so that the whole 
system of n chains will, in an enclosure at temperature 6, 
contain energy of amount nfkB, i.e.,fkO to each chain on the 
average. 

If, however, we adopt quantum views, we select the 
quantum states of motion of a chain by giving various 

integral values to the/ quantum numbers, s l9 s 2 , , <sy, 

defined by 

fp~ dq* == Si h 
r i z i A 

etc., 

where each integral is taken through a complete period of the 
corresponding co-ordinate, i.e., 2 TT/O), for q r . These con- 
ditions give in each quantum state of the chain definite 
values to the amplitudes a x , a 2 > ay, so that for example 



ELASTIC SPECTRUM OF A LINEAR LATTICE 185 

the extreme forms between which a chain can swing to and 
fro have not an " infinite variety " ; they form a discrete 
series of shapes. The energy in the state s l9 s 2 > ...... > 5 /> 

is equal to 

$i hv- -f- <9 2 hv% -f- ...... + Sj hvj, 

where v r a) r /2 TT, and in the most probable statistical 
state of the system of n chains the number which are in this 
quantum state is given by 

C exp { JJL h (s 1 v l + s 2 v 2 + ...... + s j "/)} (18.3. 1) 

where the sum of all such expressions for all sets of values of 
the integers (s l9 s 2 , ...... , s f ) is n. To find the average 

energy associated with a particular co-ordinate, say q r , we 
have to work out the sum 

C llV T 2 S r exp { jLt h (8 l V l + S 2 l/ 2 + ...... +S f V f )\ 

(18.3.2) 

and divide by the sum of the expressions (18 . 3 . 1). It is 
easy to see that (18 . 3 . 2) is equal to 

C (Es r hv r e-^Ar) [ 



where the E refers to summation of s r from to oo and the 
Z' refers to summation with regard to the / 1 numbers 
s v ...... ^r-i' s r -\- 1' ...... s f ver a ll possible values. The 

summation of (18.3. 1) can also be represented by 

C (E e'^'r) [E f exp {-ijih(s l v l + ...... + 8,^ v r ^ 



Hence the average energy sought for is 

E s r Jiv r e~* 8rhv r 



2 e- 



and this, as before, turns out to be 

Ay, 

1 ' 



186 STATISTICAL MECHANICS FOR STUDENTS 

On quantum views then the average energy of any chain 
in the system of chains is 

. . . (18.3.3). 



r-l ef^r - 1 



Since 



we can write (18.3.3) as a definite integral regarding c/2l 
as equivalent to a differential of frequency, Sv. Thus the 
energy of the chain is 

2 lr f hv , 

_ J ------ d v 

CJ #* 1 

where the upper limit is fc/2l. 

18 . 4 A Superficial Lattice. A lattice of /particles arranged 
in a square of side I with i particles in a row, and i rows 
in the whole lattice (i 2 = /), can be treated in a very similar 
way, by identifying it for practical quantum purposes with 
its limit, viz., a square elastic sheet with a mass M per unit 
area and a surface tension of T per unit length. The axes 
OX and OY lie along two sides of the sheet, and the sheet is 
supposed to be fixed along its boundary. It is known that 
if is the displacement of a point (x, y) of the sheet, it is a 
function of x, y and t which satisfies the differential equation 



T 

where c 2 = - . 

M 

From this we find a solution which satisfies the boundary 
condition, that = if x = or I or if y or I. It is 

* * 77 x . s TT y 

sin* 



i I 

where the q r8 are i 2 (or/) " co-ordinates " following a harmonic 
law of variation with time 

q rs = a n sin (o> w < - e,,). 



ELASTIC SPECTRUM OF A LINEAR LATTICE 187 

The a rs (amplitudes) and r8 (epoch-angles) are arbitrary, 
but the natural pulsations of the lattice are given by 



as can be easily verified from (18.4.1). 

All the necessary details for quantisation can be easily 
supplied by the reader from the previous section, leading to 
the average energy of any atomic sheet (regarded as a com- 
plex molecule) belonging to a system of such sheets im* 
mersed in an atmosphere of gas molecules. It is 



(18.4.2) 



r-n-i e^r* - 1 

s\ 

where v n = (r 2 + s 2 )*. 

To convert this into a definite integral in the same fashion 
as before, we observe that all the natural frequencies of a 
lattice which are not greater than a definite frequency v, 
are the same as the number of pairs of positive integers the 
sum of whose squares is not greater than (2/*>/c) 2 . If the 
integers are plotted on squared paper ruled in unit lengths, 
we can realise that this number is the number of units of 
area in a quadrant of a circle drawn on this paper having a 
radius 2lvjc ; so it is 



It follows that the number of natural frequencies of a 
lattice which lie in an elementary range v to v -f- 8v is equal 
to 

Sv .... (18.4.3) 



Thus to arrive at a definite integral which replaces (18.4.2) 
we must collect all the terms which correspond to v + $v > 
v r > v \ these will supply one element to the integral ; 
their individual value is taken as liv\(e^ v 1), and the 



188 STATISTICAL MECHANICS FOR STUDENTS 

number of them is (18 . 4 . 3). So the energy of a lattice on 
the average is 



dv (18.4.4) 

where v f = t (i 2 + i 2 ) 1 

2i i 



e? hv 1 
c 



CHAPTER XIX 

THE ELASTIC SPECTRUM OF A CUBICAL LATTICE 

19.1 Some Preliminary Mathematical Statements. 

I. The wave equation. 

If f(x, y, z, t) is a function of four variables which satisfies 
the partial differential equation 

S 2f d 2f d 2f ^ l d 2f 

Sx* dy* dz* c* dt*' 

it is called a " wave function." It can be easily verified that 
any function of the expression ax + fty + yz + f]t is a wave 
function if the constants a, /?, y, 77, satisfy the relation 

a 2 +/? 2 +y 2 -^ . . . (19.1.1) 

The reason for the name is fairly obvious in this special case. 
If x, y, z, t, represent space and time co-ordinates, the value 
of the function at a place x l9 y^ z l and time t l is the same as 
its value at the place x 2 , ?/ 2 , z 2 and time t 2 provided 

ax l + fiiji + yz l + r]t L = ax 2 + /fy 2 + yZ 2 + rjt 2 . 
In short, the value which the f unctign has on any plane 

ax + py + yz + 7]t l 
at time t 1 was or will be the value on the plane 

ax + fly + yz + ^ 2 - 

at time t 2 . These planes are parallel to one another, and 
separated by a normal distance 



i.e., according to (19 . 1 . 1) by the distance 

c ( ~~ t). 



189 



190 STATISTICAL MECHANICS FOR STUDENTS 

This clearly can be taken to refer to the propagation of any 
value of the function in a direction defined by the direction 
cosines a/ (a 2 + /? 2 + y 2 )*> etc., with a velocity c. The con- 
stancy of the value of the function over a whole plane at 
one instant is indicated by the use of the term " plane 
waves. " As the wave -equation involves only first powers 
of the differential coefficients, the sum of any number of 
particular solutions is also a solution ; so the propagation 
of any quantity by a group of plane waves in any number and 
in any directions would still yield at each point and instant, 
a quantity satisfying the wave equation. 

An important special case of this occurs when the function 
f(u) is a circular function sin u or cos u. Expansion of these 
will also yield expressions such as 

sin sin n sin sin 
ax plj yz -nt 
cos cos cos cos 

(there being sixteen of them if we ring all the changes on the 
sin and cos). These individual expressions also satisfy the 
wave equation, as can be easily seen by noting that second 
differentiation with respect to x yields the same expression 
multiplied by minus a 2 , and so on. 

II. Some remarks on the propagation of a disturbance 
through an elastic solid regarded as an ideal continuous 
medium will be necessary. The proof of the statements will 
be found in Love's Treatise on the Theory of Elasticity, or 
similar works. 

An isotropic solid has two elastic moduli ; the " bulk 
modulus " or the quotient of a uniform hydrostatic pressure 
by the fractional diminution in volume resulting from it ; 
the " modulus of rigidity," or the quotient of a tangential 
stress applied to one face of a rectangular block by the shear 
of this plane past the opposite plane (supposed fixed), 
resulting from the stress. All the elastic constants of the 
material can be expressed in terms of these two moduli, 
denoted by K and /x respectively. 

If such a medium is distorted, and we represent by , 77, , 
the components of the displacement of an element of the 
medium which was originally at a point #, y, z, then , 77, 



ELASTIC SPECTRUM OF A CUBICAL LATTICE 191 

are each functions of the three variables x, y, z* the form of 
the three functions depending on the nature of the distortion. 
The quantity 

d J -L- ^ 4- ^ 

dx + Sy + dz 

measures the " dilatation " or fractional change in volume 
of the element which is now at the point (x, y, z). We shall 
denote this by 8. This element has not only suffered (in 
addition to its general displacement) a change in size, but 
also a rotation, trr, whose amount is defined by the three 
axial components 

1 /3_ &A _ 1 /9| dj\ _ _ 1 /3, 8A 
&1 - 2 \dy ~ dzj ' Wa - 2 Vai fa/' 3 ~ 2 \dx ~ dyj' 
If instead of considering a static distortion, we consider a 
state of vibration existing in the body, then , 17, , and of 
course w^ w z , w s , d, are functions of i as well as of x, y, z. 
It is known that 6 satisfies the equation 



so that the dilatation is propagated with a velocity equal to 
j ( K + 4 fJi/3)/p J * where p is the density of the medium. 
Moreover, each component of the rotation satisfies an 
equation 



This involves a wave propagation with a velocity 

It should be realised that the components of the displace- 
ment (, 77, ) do not in general satisfy individually either of 
these wave equations. They could not, of course, satisfy 
both ; i.e., definite values of f , 77, , could not be propagated 
with two velocities. However, there are special cases of 
vibratory motion in which no rotational motion exists, 
and then the displacement-components satisfy an equa- 
tion similar to (19.1.2). There is one wave-velocity 
{(K + 4 p,/3)/p}*. It is known that for a single plane wave 

* N.B. | is in general not a function of x alone, but of all three 
variables (x t y, z), etc. 



192 STATISTICAL MECHANICS FOR STUDENTS 

propagated under such circumstances the displacement of 
an element is to and fro in a direction parallel to the direction 
of propagation of the wave. In other words a pure " dilata- 
tional " wave is longitudinal. There are also special cases 
in which no dilatation exists, the wave is purely a rotational 
one and the displacement-components then satisfy an 
equation similar to (19 . 1 . 3). Under these circumstances the 
displacement for a single plane wave takes the form of a 
vibration in the plane of the wave. Thus the rotational 
wave travels with a speed (/x/p)* and is transverse. 

The energy of strain in any element of volume is, per unit 
volume, known to be 



...... (10.1.4) 

dy dz J 

19.2 The Motion of a Cubical Lattice. In the case of a 
linear or superficial lattice we were able to appeal to the 
results for a linear or superficial continuum, knowing that 
the discrepancy between the higher frequencies of the 
lattice and the corresponding frequencies of the continuum 
would be rendered negligible in quantum applications by 
the rapidly diminishing importance of the corresponding 
average energies. We make the same kind of appeal in the 
case of a cubical lattice and for the same reason ; that is the 
justification for the rather long digression just finished. 
Without the information contained in it, any treatment of 
the lattice is apt to be very vague and not too cogent. 

The lattice will have the atoms when at rest arranged in 
cubical order so that there are i atoms in any row of length I 
parallel to an edge. Thus in a volume Z 3 there are / atoms 
where/ = i 3 . Any atom will be identified by three positive 
integers r, s, u ; f being its number in a row parallel to OX, 
5 in a row parallel to OY, u in a row parallel to OZ. The 
suffix (rsu) will refer to this particle, and we can by methods 
similar to those used for a linear lattice express each com- 
ponent of the displacement of the atom (rsu) thus 



ELASTIC SPECTRUM OF A CUBICAL LATTICE 193 

f * * ^ r TT x s TT y u TT z 

&, u = S Z Z Q wu sin -y- sin -y- sm j + 

r-U-lu-l III 

and similar expressions for rj rgu9 rsu where the + written at 
the end is to indicate that we can have seven more series if 
we like which involve terms such as cos r<{> sin sift sin u\, 
cos r<f> cos sift sin ux, cos r0 cos sift cos u\, etc., obtained by 
ringing the changes between sin and cos. (^, ift 9 x are written 
for 77 x/l, etc., for convenience.) The Q, rsu are generalised 
co-ordinates, and a little thought will show that if we used 
all the eight series suggested there would be eight of them 
for a particular set of values of r, s, u. As there are three 
components of displacement, this would give us twenty -four 
co-ordinates for each set of (r, s, u) values. In consequence, 
when counting over all the terms of each series we should 
have 24 i 3 or 24 / co-ordinates in all. But this is just eight 
times as many as are required to co-ordinate / particles each 
with three degrees of freedom. We clearly must limit the 
terms in the series in some way. In dealing with the linear 
and superficial lattices, we only retained sine terms, since 
only with them could we satisfy the conditions at the 
boundaries. The matter is not so simple here, but the clue 
to a successful reduction of terms lies in suitable boundary 
conditions, and in the fact that the expressions for , 77, , 
must be separable into terms expressing the dilatational 
motion and terms expressing the rotational. We shall 
write down the following and justify them presently. 

. __ L t, L f #! cos rcf) sin sift sin u\ 

7=i =iu-i 1 + Q'I sin r cos s ^ cos 



= X Z X . , 1(19.2.1) 

y ' 



* . , 

+q 2 cosr(f> sm sift cos 



+ 2'a cos ^cos sift sin u\ 

where the six quantities q l9 ...... , q'% have in general a 

different set of values for each choice of the integers r, s, u. 
In fact they should really be written as (qi) r8Uf etc., but to 
avoid clumsiness, this complete suffixing is not used, since 
no ambiguity will arise on that account. There are 6/ 

8.M. 



194 STATISTICAL MECHANICS FOR STUDENTS 

co-ordinates still involved, but assuming that q l9 q 2 > q$ refer 
solely to the dilatational motion and q\, q' 2 , q' 3 to the 
rotational, we arrive at three conditions which must then 
hold for each set of six, and this reduces them effectively to 
three independent co-ordinates. To see this, let us work out 
the expressions for the dilatation and the components of 
rotation. 

6 = - S S Z ~~ ( rq ~*~ 5?2 "*" Uq sin r sin 5 sin 



J 2 + U( l 3) cos T 9 cos 5 r cos 

.. r _, , V o<j 3 u^2/ s i n r< cos 5l A cos W X 
uJi - v^L^/^i / / / v / . . 

1 / ( (sq 8 w? 2 ) cos r<f> sm 5^r sm u\ 

and two similar results. 

If the 3/ co-ordinates g'j, q' 2 , q' 3 are only to supply the 
rotational part of the motion, then we must have a relation 

rq' l +sq' 2 +uq' 3 = Q . . (19.2.2) 

for each set of values of r, 5, u. If q l9 q 2 , q 3 are only to supply 
the dilatational part, then we must have the two relations 

= = -. . . . (19.2.3) 
r s u v 

for each set of values of r, s, u. It will be seen that with 
these relations, the dilatation vanishes at the boundaries 
where sin <f> or sin */r or sin x is zero ; also the rotation at any 
point of one of the boundary faces is around a line normal 
to the face. To sum up, we can, putting q^ = rq, q 2 = sq, 
q 3 = uq, write instead of (19 . 2 . 1) 

If sin m } 



. (19.2.4) 
sm T 9 cos s v cos U X J 

and two similar equations, where the relation 

r ?'i + ^2 + ttff'a = 

holds, so that we have reduced the co-ordinates to three 
independent co-ordinates, q and any two of q' l9 q' 2 , q'$> for 
each set of values for r, s, u. 
The dilatation is then given by 

= - 2 2 Z {(r 2 + 5 2 + u*) q sin r<f> sin s$ sin u\\ 

1 (19.2.5) 



ELASTIC SPECTRUM OF A CUBICAL LATTICE 195 
and the rotation-components by 

2w l = - E Z Z {(sq' 3 - w?' 2 ) cos r<f> sin sift sin ux} 

P (19.2.6) 

and two similar expressions. 

This is a matter of geometry and analysis. In a static 
distortion, the q and q' quantities are constants ; but when 
vibration takes place they are functions of t, and we appeal 
to dynamics to discover the forms of the functions. A glance 
at (19.1. 2) and (19.2. 5) shows that 

q = <*>r*u cos K t m ) 
where 



or oj r8U = (r 2 + s 2 + u 2 ) 1 . . (19.2.7) 

c being the velocity of a longitudinal wave through the body. 
Similarly the first of (19 . 1 . 3) and the first of (19.2.6) 
show that 

Sq'z u q*2 ^ COS (^'rsu t -~ 'rsu) 

where 

a>' m = ^ (r 2 + s 2 + u*)* . . (19.2.8) 
l 

c' being the velocity of a transverse wave. Two similar pro- 
portionalities for uq\ rq'a and rq' 2 sq\ show that 

q'l == a 'r*u COS (o/ r , u ^ 6 'ru) 

q' %~ b' rsu cos (co' raM ^ f rsu ) 
2'a = c/ u cos (a)' r8U t c' nu ) 
where 

r a' rsu + s b' rm + u c' rm = 0. 

19 . 3 The Energy of the Lattice. The reader will probably 
be anticipating the direction in which the argument is going, 
but to complete it we must form the expression for the 
energy and show that it involve^ only squares of the g, q' and 
the q, q'. This may appear to be an appalling task in view 
of the complexity of the expressions obtained. Fortunately 

02 



196 STATISTICAL MECHANICS FOR STUDENTS 

it is easy to show that an enormous cancellation of terms 
takes place automatically. 

Let us deal with the kinetic energy first. This is obtained 
by putting dots over the q and q' in (19 . 2 . 4), then squaring, 
multiplying by P dx dy dz, and integrating between the 
limits o and / for x, y, z. Now in such integrations all 
integrands which involve sines or cosines of different multiples 
of (f> (i.e., 77 x/l), or of 0, or of x> add nothing to the result, 
since 

r i 

I . r TT x . r f TT x j . .- , f 
sm - sin - dx = if r r 



r TT x r' 77 x , .. j t 

cos cos - dx = if r =f= r 



I T 77 X T 77 X 

and sin - cos -* dx = in any event. 

%j 

This wipes out a great body of terms, and it follows that 
only the squares of individual expressions such as are written 
within the brackets in (19.2.4), will yield finite amounts. 
Even in these we shall find in the end no products such as 
qq\, q q'& q ?' 3 , since in the case of such products we are 
involved in integrations such as 

sin r<f> cos rcf> sin sift cos s0 sin ux cos ux dx dy dz, 

and these lead to a sero result. In the end a little inspection 
will show that the kinetic energy is the sum of 



, 

O 

over all the sets of values for r, s, u. 




ELASTIC SPECTRUM OF A CUBICAL LATTICE 197 

Turning now to the potential energy, we glance back at 
(19.1.4), which has to be multiplied by dx dy dz and inte- 
grated throughout the body. The term (< -f- 4 ju/3) 2 , in 
view of the remarks just made on the ndture of the inte- 
grations, clearly only involves the squares of the q co- 
ordinates ; the term 4 ^(w-f + w 2 2 + m 3 2 ) involves the 
squares of ' (<s#' 3 ^'2) an d such like quantities. The 
quantity within the brackets { } in (19.1.4) looks as if it 
might give trouble. As a matter of fact, if the reader likes 
to take the trouble of working it out in detail, he will find 
that the integral of it vanishes, and if he shirks the task he 
can take the statement on faith. Still it looks as if the 
terms referred to in the previous sentence might give 
products ; for they will yield on integration the sum of 
expressions such as 



But the expression inside the bracket is equal to 



in view of (19 . 2 . 2). Thus the statement that the energy of 
the lattice involves only squared terms in the generalised 
co-ordinates and velocities is justified. 

So we arrive finally at a conclusion much the same as in 
the previous chapter ; a lattice such as we have described 
of volume Z 3 , and subject to the defined boundary conditions, 
can be regarded a large complex molecule with 3/ internal 
degrees of freedom which, as far as the energy expressions 
are concerned, is analogous to a molecule with a number of 
internal oscillators. In this analogous molecule, / of the 
oscillators have each one degree of freedom, and each one 
has a period belonging to the series determined by a law such 
as (19 . 2 . 7). That one for instance which corresponds as 
regards period to the integers r, s, u, has a direction of 
motion whose direction cosines are proportional to r, s, u. 
(See equation (19.2.3).) Carefully note that these remarks 
are not made concerning the atoms of the lattice, but about 
the oscillators of a hypothetical molecule whose energy 



198 STATISTICAL MECHANICS FOR STUDENTS 

function is formally identical with the formula for the energy 
of the lattice. There are / other oscillators in the molecule 
which have each two degrees of freedom, and each one ,a 
period drawn from a series determined by ( 1 9 . 2 . 8) . The one 
which corresponds to the integers r, $, u vibrates in a plane 
whose normal has direction cosines proportional to r, s, u. 
(See equation (19.2.2).) 

19 . 4 The Statistics of a System of Cubical Lattices. 
Conceive then a- system of n such cubical lattices capable of 
interchanging energy one with another, via, for example, 
the medium of an assemblage of gas molecules. We can 
derive momenta for each q co-ordinate of the lattice by the 
usual differentiation of the kinetic energy partially with 
respect to each q velocity ; set up the usual machinery of a 
phase-diagram of 6/ dimensions, partition the representative 
points, and count the complexions in each statistical state. 
Then introducing the hypothesis of quantum states, we 
arrive at the conclusion that in the most probable state the 
number of lattices in the system of lattices which are in a 
quantum state, defined by the condition that the action- 
integrals of pdq (each one integrated throughout one period 
of the particular q) shall be certain integral multiples of h, 
are proportional to 

e -n 

where e is the energy of that state and is equal to 

Z Z 2 [a rm Uv nu + 2 b rltt hv' n \ . .(19.4.1) 

f-1 *-l M-l [ J 

the a r8H , b r8U being the 2/ quantum numbers of the particular 
quantum state considered. The factor 2 in the second term 
arises from the two degrees of freedom of one set of oscillators 
in the analogous molecule. The average energy of a lattice 
is obtained as usual by evaluating 



where the 2 in the numerator and denominator indicates 
a multiple summation for all values from to <# of each one 
of the 2/ quantum numbers a rsu , b r8U . The symbolism has 



ELASTIC SPECTRUM OF A CUBICAL LATTICE 199 

become somewhat cumbersome, and it will enable the 
reader to arrive at the result more easily if we temporarily 
simplify the notation and consider an expression such as 






S E 2 . . . (ax + by + cz + . . .) 



2 2 S 
a=0 6 = c=0 



The letters x, y, z . . . replace for the moment the individual 
hv, while a, 6, c, ... replace the quantum numbers. 

Now the denominator of this expression can, as a little 
thought will show, be factorised. It is equal to 



(,!/"") If 



The numerator can first of all be separated into a number 
of expressions such as 



oo oo oo 



6=0 c=0 



and this part of the numerator can as before be factorised 
into 



( S ae- a x ] ( E e~ b y \ ( E e 

\a = / \6=0 / \c=0 



When this part of the numerator is divided by the denomi- 
nator the result is 



a-0 

and this is 



x 



g^ J 

(See section 14 . 3). 
Thus we see that the original expression is equal to 

* -i y + z + . 



200 STATISTICAL MECHANICS FOR STUDENTS 

In this way it can be established that the average energy of 
a lattice, viz. (19.4. 2), is equal to 

S E Z\ . - hv ______ +2 ____ hl/rsu _ 1(19.4.3). 

r-i f .i u =i [ exp (phv r8U ) - 1 exp (^hv' T8U ) 1 J 



We can convert (19.4.3) into a definite integral just as in 
previous cases. The natural frequencies of a lattice for the 
q co-ordinates, which are beneath a certain value v are in 
number the same as the number of positive integers the 
sum of whose squares is equal to (2 lv/c) 2 . (See equation 
(19.2.7).) This is the same as the number of points dotted 
at the corners of unit cubes inside one octant of a sphere 
whose radius is 2 Ivjc ; and so it is just one -eighth of the 
volume of this sphere, i.e., 



77 /2 lv\ 

6 \ c I 



2 lv\ 3 4 77 

or - 

3c 



Thus the number of frequencies which lie in a range of fre- 
quency v to v + Si> is 

477Z 3 2 , 

- - v*8v, 
c 3 

and so we convert (19.4.2) into a definite integral by 
collecting terms which satisfy 

v + Sv > v nu > v 

t i ' -^ ' --^ / 
v + bv > v nu > v 

to supply one element of the integral giving each of these 
terms the value which corresponds to v and v and multi- 
plying by the number of terms, viz. (4 TT Z 3 /c 3 ) v 2 8v and 
(4 TT Z 3 /c 3 ) i/' 2 Si/. We can obviously in the final result drop 
the stroke over the v and write the average energy of the 
lattice of volume Z 3 or V to be 



1 2 
- + - 

3 c 



2 \ 4 77 Av 3 , 

; ) - v ---- d*> . . .(19.4. 4). 

/3 / e^" 1 v ; 



19 . 5 Standing Waves. The boundary conditions which 
were imposed so as to give a definite character to the 
problem in hand can be exhibited in another light. Reverting 
for a moment to the linear lattice, it will be observed that, 



ELASTIC SPECTRUM OF A CUBICAL LATTICE 201 

selecting any term from the series which expressed the dis- 
placement, say a r sin r TT x/l . cos a> r t, it can be written 

1 . T TT , ,v , 1 T 7T , , ,v 

- a r sm ( x ct) + - a r sin _ (x + ct). 

If we therefore considered an unlimited string and two 
progressive waves 

> 1 . r TT 



J 

and 



. ,. 
r (x - ct) 



r T* i i 

-- a r sm (x -f 



passing along it, the first in the direction of x positive, the 
latter in the opposite direction, the result would be that the 
appearance of progression is absent. The string would 
behave as if it were divided into segments of tength Z, each 
segment vibrating as, an individual in its r th upper partial. 
We would have a group of " standing waves." Thus the 
most general vibration can be regarded as a composition of 
standing waves of different periods each wave being equi- 
valent to the existence of two progressive waves of equal 
amplitude travelling in opposite directions. 

The superficial lattice exhibits the same features. The 
expression for one harmonic of the complete vibration can 
be regarded as the sum of four progressive waves such as 

f = a n cos K -~ (ax + fy ct) 
4 l 



L 



= 7 a rs COS K ? ( aX + fy + Ct ) 



- - a n cos *L? (ax ft/ ct) 

= -a n cos --^ (ax fy + ct) 
4 I 



where 

r 

(r* + 

K = (t 



202 STATISTICAL MECHANICS FOR STUDENTS 

The first two represent two oppositely propagated pro- 
gressive waves, one in the direction (a, j8), one in direction 
(a, /?) ; the second two represent waves in directions 
(a, /?) and (a, j9). All four have the same speed c. 
(Note that -^ a, -[- j8 are direction cosines.) Thus the lattice 
with fixed edges might be regarded as a part of an infinite 
lattice with four sets of widely-extended waves for each 
(r, s) harmonic travelling across its surface. 

Taking now a special term from the complete expression 
for a cubical lattice, say 

r 77 x s 77 y u TT z 
cos ~- sin -~~ sin ~~ 



r TT x s TT y u TT z 
= r a rsu cos j~ sin y- sin j cos a) rsu t 

it will be found that this can be regarded as the sum of eight 
progressive waves. (We drop the suffixes as unnecessary at 
the moment.) 

1 K7T / \ 

= - ra cos -y I ax + py +yz ctj 

1 K7T ( \ 

% = - ra cos -y- (ax + fy + yz + ctj 

> ! K7r f \ 
f = - ra cos -y- [ax py yz cti 

> ! ^^^ \ 

f = - ra cos -y- (ax py yz + ctl 

f = - ra cos y f aa; + py yz ctj 

1 /C7T/ n \ 

= -g ra cos -y I au; -f- ^y yz + ctj 



1 Krrf \ 

= -ra cos -y- ^ax jfy + y^ cM 

= - ra cos ^ ^aa? j8y + y + cM 



ELASTIC SPECTRUM OF A CUBICAL LATTICE 203 

where a = (f8 + J 2 + ^ etc. 

K = (r 2 + s 2 + u 2 )*. 

Each of these represents a progressive wave of amplitude 
ra and speed c travelling in one of the eight directions 
given by ringing the changes on the signs of the direction 
cosines i a, ^ /J, ^ y. In the same way the vibration- 
component of 77 can be split up into progressive waves of 
amplitude ^ sa, having the same harmonic factors, and 
likewise for . When we associate the whole twenty -four 
into eight sets of three, each trio of course involving , 77, , 
it will be found that the resultant amplitude has the same 
direction as the corresponding direction cosines of the trio. 
In short, the waves arc longitudinal. The transverse can be 
treated in a similar manner. 

Now the special feature referred to at the beginning of this 
section is this : Taking a point o, y' ', z f on a face of the cube 
and the point I, y' ', z' on the opposite face directly opposing 
it, we see that for the longitudinal standing wave 

rau = ra cos r<f> sin sift sin ^x cos (a> nu t nu ) 

etc. 

the displacement components rj r8U9 rsw vanish while the 
individual harmonic constituents of g rsu are 

S TT y' U TT Z' / \ 

in j sin j cos I a) nu t rsu l at o, y , z 

s TT y' u TT z' ( \ 

ra sin y sin j cos I a> rsu t * rsu J at I, y', z'. 

Each of these harmonic vibrations has the same amplitude 
ra sin (s TT y f jl) sin (u TT z'/l), and is in the same or opposite 
phase. These remarks can be extended to the transverse 
waves. 

Thus we find that the behaviour of the cubical lattice 
simulates that of a cubical portion of an infinitely extended 
lattice through which are travelling a specially selected 
group of progressive waves, one half longitudinal, one half 
transverse. These waves produce standing waves in such a 
manner that at points directly facing each other on two 



ra sin 



204 STATISTICAL MECHANICS FOR STUDENTS 

parallel faces of the cube, the normal components of the 
longitudinal vibration are always in the same or opposite 
phases, while the components in the surface of transverse 
vibration have the same character. 

The drawback of this analysis of the vibrations of a lattice 
(which is originally due to Jeans) for the problem in hand is 
that it really concerns a continuous medium. The number 
of components is really infinite, whereas for our purpose we 
have to cut the series off sharp in an artificial manner since 
we are thinking of a group of discrete particles. Neverthe- 
less the quantum result (19 . 4 . 3) or (19 . 4 . 4) obtained from 
it, was used by Debye towards the close of the first period 
of quantum history to deal with the question of the specific 
heat of solids and its success was so marked that when a 
little later Born and Karman actually solved the difficult 
problem of the real periods of a lattice proper, the improve- 
ment in the agreement of calculation with observation was 
not so noticeable. Born and Karman's analysis is too long 
and difficult to reproduce here, but a reader with rather more 
than the usual run of mathematics at his command could 
consult Born's works, Dynamik der Kristallgitter , or Theorie 
des Festen Zustanden. As stated, the analysis for a con- 
tinuous medium was made by Jeans in 1906 in connection 
with the problem of full radiation, following up an earlier 
hint of the late Lord Rayleigh's. The ether being regarded, 
in the familiar fashion of those days, as a continuous medium 
capable of propagating transverse vibrations alone, the 
energy of radiation in a volume Z 3 of it could be determined 
by multiplying the average energy corresponding to a 
frequency by the number of natural frequencies in a range 
v to v -f- Si/, viz., 8 77 I 3 v 2 Sy/c 3 ,* where c is the velocity of 
light, and summing over all the frequencies. We have seen 
that statistical mechanics distributes the energy on the 
average between these frequencies in the same manner as 
it distributes it between the ordinary co-ordinates and 
momenta of a particle system, because the motion can be 

* The number of longitudinal frequencies (47rZV5v/c 3 ) is zero since c 
for longitudinal vibrations in the ether is infinite, as this was considered to 
be an incompressible medium. 




ELASTIC SPECTRUM OF A CUBICAL LATTICE 205 

analysed in terms of specially chosen co-ordinates and 
momenta (in a wider sense), and to each co-ordinate there 
is one corresponding frequency. This procedure has, of 
course, still to be justified to the reader, but all in good time. 
If then we adopt classical views, we ascribe kd to each 
frequency, and find for the energy per unit volume of full 
radiation in the ether (" full " in the sense that it is in 
statistical equilibrium with its surroundings) 



. - .< 



This is the famous Rayleigh- Jeans Law. Apart from the 
difficulty that the integral is infinite in value if we actually 
go to the upper limit, as we strictly should do if we regard 
the ether as a continuous medium, it does not fit the experi- 
mental facts within any finite range of frequencies except 
when well down in the infra-red. It concentrates all the 
energy as it were in the upper frequencies. Even if we 
artificially reduce the upper limit from oo to the highest 
known frequencies (say y rays), and so evade the difficulty 
just mentioned, it is clear that as between the relatively 
few (though on our normal ideas of counting, enormous) 
degrees of freedom say in a cubic inch of iron and the 
relatively great number in a cubic inch of ether, the energy 
in a state of equilibrium should nearly all be in the ether 
and the iron just a trifle above absolute zero. This was 
a serious difficulty for older views, even before experi- 
ments on the spectral distribution of energy in full radiation 
had reached such precision as to pronounce a definite 
verdict against (19 . 5 . 1) in detail. These difficulties dis- 
appear if we replace kd by the quantum result for the 
average energy which is not the same for all frequencies, 
but diminishes with frequency in a marked way on account 
of the exponential term in the denominator. We then 
obtain for the energy-density of full radiation 

8 TT h f v 3 

TT-TTm \dv . . . (19.5.2). 

exp (hv/k6) 1 v ' 




206 STATISTICAL MECHANICS FOR STUDENTS 

This is Planck's famous expression, although he obtained 
it originally in a very different way (actually five or six years 
before Jean's analysis). Of course (19 . 5 . 2) is a particular 
case of (19.4.4) with the longitudinal velocity infinite and 
the transverse velocity (here written c) put equal to the 
velocity of light. 



CHAPTER XX 

THE SPECIFIC HEATS OF SOLID BODIES 

20 . 1 Einstein's Theory for Monatomie Solids. According 
to the classical statistical theory, the specific heat per gram- 
atom, or atomic heat, of all monatomic solids should be 
3R (5-94 calories per degree) at all temperatures. This, by 
the way, is the atomic heat at constant volume while the 
specific heat which is usually measured is obtained under 
conditions of constant pressure and is somewhat larger than 
the former, the difference being given by the formula 



where a is the coefficient of thermal expansion, K the iso- 
thermal compressibility, and V the atomic volume. The 
theoretical s p will, of course, vary somewhat from substance 
to substance by reason of differences in atomic weight and 
thermal and elastic properties, but the values are round 
about 6'4. Some metals keep reasonably near to this at 
ordinary temperatures, e.g., silver, copper, lead, aluminium 
and zinc. On the other hand, substances like boron, beryl- 
lium, silicon, carbon, have values much too small at ordinary 
temperatures; diamond, for example, has an atomic heat 
0-75 at 50 C., and in any case all the substances show a 
marked decrease of specific heat with falling temperature. 

Einstein was the first to suggest that the explanation of 
these facts lay in an abandonment of the equipartition 
theorem. He illustrated his suggestion in a rough and ready 
way by treating each atom in a monatomic solid as a har- 
monic oscillator with a frequency v and three degrees of 
freedom. Thus the heat energy of a solid with / atoms 
would be 




208 STATISTICAL MECHANICS FOR STUDENTS 

and the thermal capacity would be equal to the differential 
coefficient of this with respect to 0, i.e., to 



where 

hv 

x =,*=-. 

This only approaches the classical value 3R as x approaches 
zero, or 6 approaches infinity ; for 



On the other hand, the expression diminishes indefinitely as 
x increases to an infinite value, i.e., as 6 approaches zero. 

About the same time several attempts were made to evade 
the difficulty along classical lines. These took the form 
of assuming that with falling temperature some of the 
degrees of freedom become " frozen-in," to use a picturesque 
phrase, i.e., that certain linkages otherwise free, and per- 
mitting some relative movement of different parts, become 
completely rigid. The trouble about these suggestions lay 
in the fact that if this were the case, the bodies should be 
much more difficult to compress at low temperatures than 
at high, and this is not so. Further, some time later, Einstein 
was able to derive a value for his mean frequency v in terms 
of the atomic weight, density and compressibility of the 
material, and it turned out to be of the right order of magni- 
tude to suit his specific heat formula. To be sure his assump- 
tion of one mean frequency was bound to produce a formula 
not completely in line with the facts, although a decided 
improvement on Dulong and Petit's law. For one thing, 
experimental work inspired by his result soon discovered 
that his value for s v fell far too rapidly with temperature, on 
account of the exponential term in the denominator. In 
1912 Debye published a very long paper on the whole matter. 
The first part consists of an investigation of the natural 
frequencies of a solid continuum leading with the help of the 



SPECIFIC HEATS OP SOLID BODIES 209 

quantum theory to the results of the previous chapter. He 
then proceeds to apply them to a lattice for which the fre- 
quencies must have an upper limit. It is in this step that 
the weak spot of Debye's method lies, yet his final results 
constitute as great an advance beyond Einstein's formula 
as the latter was beyond the law of Dulong and Petit. 

We saw in section (19.4) that the number of natural fre- 
quencies for a cube of volume V which lie in the range v to 
v + Sv is 

2\ 
+ pij * 2 8"- 

Hence, if v m is the upper limit, the integral of this expression 
from o to v m must be the whole number of frequencies, i.e., 
3/, where / is the number of atoms in volume V. Thus 



4.V/1 2\ 
7 3Vc 3+ c'V 



l V,-3JO -81 

4 TT V (c 3 + 2 c 3 ) 

and since c and c' depend on the density and elastic constants 
of the substance, v, m is calculable in terms of these quantities. 
The heat energy of the body is by (19 . 4 . 4) 



(c 



hv* , 

dv 



which is equal to 



_ 9fk0 t* 



J e-\ 



dx (20.1.2) 



where we write x for hv/kd and x m for hv m /kd. The integral 
is, of course, a function of x m , say </r (x m ) 9 so that the heat 
energy at the temperature 6 is 

W . tM. 

x m 

S.M. P 



210 STATISTICAL MECHANICS FOR STUDENTS 

The differential coefficient of this with respect to is the 
thermal capacity of / atoms. This coefficient is 



3 (a*) -^-j- r v~m/ ^ i 3 ^ d 
_ /9/A 27/fe0 



3; 3 g^ 1 

.... (20.1.3). 

0* I P 1 P^Wl\ 1 

m J o 

The expression in (20 .1.3) is a definite function of x m . 
Thus Debye arrived at the conclusion, that the atomic heat 
for any monatomic solid is represented by a universal 
function of a pure number x m which has a characteristic 
value for each solid at a given .temperature. This number 
is the ratio of the quantum of energy for the maximum 
frequency of the solid to the average energy (kinetic + 
potential) at the temperature per degree of freedom on the 
equipartition law. Obviously hv m /k has the physical 
dimensions of temperature and if we call hv m /k the " charac- 
teristic temperature " of a solid whose maximum frequency 
is v m , then x m = 0J0, where 9 C is this characteristic tempera- 
ture. Thus the atomic heat can be written according to 
Debye as D (OJ0), where D (y) is a function of the variable 
y defined by 



If y is large, the second term in (20 .1.4) becomes 
negligible, and since it is known that 



it follows that for large values of y, the function D (y) 
approximates to 

12 7T 4 R 

Hence by (20 . 1 . 4) the value at low temperatures of 8 9 is 

12 77 4 R 

"TV"'* 



1 SPECIFIC HEATS OF SOLID BODIES 211 

and so the specific heat varies as the cube of the temperature 
when it approaches absolute zero. This is in excellent 
agreement with the experimental facts. We would expect 
the approximations at low temperatures to be good ; for in 
such a condition all the energy practically resides in those 
periodic motions which have the very smallest frequencies 
possible to the structure. The wavelengths in the solid for 
such frequencies are then so much longer than the separation 
between molecules that the assumption of a continuous 
medium inherent in certain parts of Debye's reasoning is 
practically justified in such extreme conditions. 

If on the other hand y is small enough to permit of the 
approximation, e y = 1 + y, the function D (y) approxi- 
mates to 

y- f x 2 dx 9R, 

or 3R, and so we find that the specific heat at high tempera- 
tures has 3R as a limit, as it should be. 

The result (20 .1.4) has been tested for about half-a-dozen 
metals in a very searching manner. There is an expansion 
of (e x 1 )~ 1 as a series of powers of x which is well known to 
the pure mathematician, and whose coefficients have been 
calculated by him as a matter of interest in other con- 
nections than the needs of physical science.* This fortunate 
circumstance enables D (y) to be calculated with not too 
much trouble for values of y ranging from small magnitudes, 
such as 0-1 to 10 or 20, or inversely to infer a value of y from 
a value of D (y). From the nature of the result if the 
specific heat of a solid A at a temperature 6 is the same as 
that of a solid B at temperature 0' ', then 6 bears to the 
characteristic temperature of A the same ratio as 6' bears to 
the characteristic temperature of B. Thus from the experi- 
mental values of the specific heats of a solid we should find 
consistent values for its characteristic temperature. But 
(20 . 1 . 1) permits us to calculate v m , and therefore C , which is 
hv m /k 9 in terms of the speeds c and c', i.e., in terms of the 
elastic constants of the solid. A comparison of the values 

* See Chrystal's Algebra, Vol. II.. Chap. XXVIII. 

P2 



212 STATISTICAL MECHANICS FOR STUDENTS 

of e determined by such diverse methods provides a very 
clear test for the general validity of the ideas involved in 
Debye's analysis. Here are some results. The values in 
Column I. are derived from specific heat data, those in II. 
from measurements of the elastic constants. 

I. II. 

Lead . . .95 ... 73 

Cadmium . .168 ... 174 

Silver . . . 215 ... 214 

Copper . . 309 ... 332 

Aluminium . . 396 ... 402 

Iron . . . 453 ... 484 

Since the atomic heat of a simple substance is f(0/0 c ), 
where / is the same function for all the substances and 6 C is 
a constant characteristic of each substance, it should be 
possible to represent the variation of the atomic heat at 
constant volume with temperature for all substances on the 
same curve, provided the atomic heats are plotted as ordi- 
nates and the values of 6/0 c as abscissae. Schrodinger has 
shown that with a suitable choice of 9 C for each substance 
this deduction is very thoroughly verified by experiment. 
For details the reader should consult Heat and Thermo- 
dynamics, by J. K. Roberts, Chap. VII. 

Diamond offers a very interesting illustration of this 
theory. On account of its elastic properties and its small 
density the velocities c and c' in the formula (20 .1.1) are 
exceptionally large. This means that the value of v m9 and 
therefore of C ; is also exceptionally great. In fact, C is 
1,860. Thus the temperature of diamond should attain an 
abnormally high value before the value of its atomic heat 
comes anywhere near to the Dulong-Petit limit. This 
peculiar behaviour of diamond had been a puzzle to the 
physicist for many years until the Quantum theory cleared 
up the mystery. 



CHAPTER XXI 

THE ENTKOPY CONSTANT OF A GAS 

21.1 Nernst's Heat Theorem. We are about to return to 
the question raised at the end of Chapter VII., concerning 
an absolute value for the entropy-constant of a gas, but 
postponed for further discussion until the modifications 
introduced into statistical-mechanical theory by quantum 
considerations had been explained. The matter is one which 
cannot be dissociated from the extremely valuable sug- 
gestion concerning entropy first made by Nernst in 1906 ; 
and no apology is therefore needed for a digression at this 
point into purely thermodynamic theory placing before the 
reader in general terms what that innovation amounted to. 

In a chemical reaction or physical change which takes 
place at constant volume and temperature, there is a simple 
connection between the heat absorbed or evolved by the 
system and the change in its free energy. Choosing 0, v to 
represent the thermodynamic variables, temperature and 
volume, and ctj, a 2 , ...... to represent other variables 

which are involved in the reaction or change, let us denote 
the internal energy, free energy and entropy by the func- 
tional symbols 

E (0, v, a l9 a 2 , ...... ), F (0, v, 04, a 2 , ...... ), 

S (0, V,a l9 a 2 , ...... ) 

respectively. The amount of heat supplied to the system in 
an elementary change is given by 



where p is the pressure ; p is, of course, a function of 
and v given by the characteristic equation of the system. 
Since E = F + S, it follows that 



00 ov oa r 

213 



214 STATISTICAL MECHANICS FOR STUDENTS 

and if the reaction or change occurs at constant temperature 
and volume, then 

8Q = 27 !? 80, + ? -I? ** 

oa r oa r 

r aF * or a f^\* 

= L - oa r VZ I - I oa r 

da r r da r \d0J r 

since S = 9F/90. Thus we obtain 



Hence the heat absorbed by the system during a finite 
reaction or change occurring at constant volume and tem- 
perature in which the a r variables vary in value from a' r 
to a" r is given by the expression 




Since the first term in this expression is the change in the 
free energy, we have here the familiar equation of the text- 
books on physical chemistry 

U=A-*H . . . (21.1.1) 

where U is the heat evolved during a reaction or change and 
A is the loss of free energy, 3 A/90 being, of course, estimated 
at constant volume. 

Now it is with the free energy that the physical chemist is 
vitally concerned. He can measure the quantity U, and so 
the simple differential equation (21 . 1 . 1) enables him to 
calculate the loss of free energy apart from an arbitrary 
constant of integration. There is no clue to the value of this 
constant on grounds of pure mathematics. Its value must 
be obtained as the result of a suitable physical hypothesis. 
Nernst suggested that trial should be made with the con- 
dition that 3 A/90 diminishes in value to zero as a limit as 
the temperature approaches absolute zero, i.e., 

3A 

->oas0->o. . . . (21.1.2). 



ENTROPY CONSTANT OF A GAS 215 

The suggestion has been abundantly supported by experi- 
mental results,* and so the equation (21.1. 2) is the mathe- 
matical form of Nernst's heat theorem, and by it is deter- 
mined the " chemical constant " of any reaction and the 
ambiguity in the solution of (21 . 1 . 1) removed. 

But from what just precedes, we can clearly write the 
condition (21 . 1 . 2) in the form 

9 F (0, V, a' v a' 2 ...... ) 8 F (fl, v, a* l9 a%, ...... ) 



if = o ; or 

S (M,a'i,a' 2 , ...... ) - S(0 9 V,a* l9 a" 2 , ...... ) (21 . 1 . 3) 

if r= o. In words, the difference between the entropy of 
the system before and after a reaction gradually disappears 
as the temperature at which the reaction occurs is reduced to 
the absolute zero as a limit. 

21.2 The Entropy-Constant and the Vapour-Pressure- 
Constant. As we saw in Chapter VII. the entropy of a 
gram-molecule of a perfect gas containing one type of mole- 
cule is 

s p log9 -Rlogp + K . . . (21.2.1) 

where * is a constant of integration, which is of no import- 
ance in dealing with the change of entropy between two 
states of the gas, but assumes quite a new significance when 
Nernst's theorem is considered in connection with it. Thus 
let us assume that at high enough temperature the vapour 
of a liquid is in such a condition that (21 . 2 . 1) is a good 
enough approximation to its entropy at temperature 9. At 
the same temperature let </> (0) be the entropy of the liquid ; 
then 



where s is the specific heat of the liquid. Of course in <j> (0) 
must also occur an undetermined constant of integration ; 

* See System of Physical Chemistry, Vol. II., Chapter XIII. ; W. C. 
McC. Lewis. 



216 STATISTICAL MECHANICS FOR STUDENTS 

but this constant is clearly settled by the value we choose 
for K in (21 . 2 . 1) ; for 

#(0)+==a,log0--Rlogi> + jc . . (21.2,2) 

since the two entropies must differ by L/0, L being the heat 
of vaporisation. 

Let us turn for a moment to the well-known thermo- 
dynamic relations which hold for the change of state from 
liquid to vapour, and which are associated with the name of 
Clapeyron. They are 

d L 



dL 



In the second, s l is the specific heat of the saturated vapour. 
Now in the condition of saturation at temperature 6 + 80, 
let the pressure be p + Sp, and the specific volume be 
v l + 8v v Conceive the vapour to expand to a specific 
volume v l + ^i + S'^u so that its pressure is restored to 
the original value p, the temperature still being 6 + 0, so 
that the vapour is unsaturated. If the temperature be so 
high that the conditions approximate to those of an ideal 
gas, then by Boyle's law 

(P + 8p) K + 8*1) = P(Vi + 8*1 + 8'*i), 
or practically 

v 8p p S'v l 

Now from the definition of s l it follows that 

8't> 
^=i+P^- 

Therefore 



Since v 2 is small compared to v l9 this can be practically 
written 



ENTROPY CONSTANT OF A GAS 217 

dp 
S P =S! + ( Vl - *,) ^ 

= i+ j by (21. 2. 3). 

Hence the second of the relations above becomes 
dL 



where s, the ordinary specific of the liquid can be regarded 
as practically equal to s 2 . Thus (21.2.5) can be regarded as 
valid at sufficiently high temperatures. But the first relation 
(21.2.3) can under these circumstances be written 

dp pL 



d log p L 

- 



To connect these two approximate results (21 .2.5) and 
(21.2.6) with our previous considerations of entropy, we 
must integrate them. Thus (21.2.5) yields 

f 
L = L + s p O \s (x) dx, 

Jo 

where L is a constant of integration which may be called 
the " latent heat at absolute zero " without necessarily 
implying any physical reality for a change of state actually 
at this temperature. Of course s p is a constant approxi- 
mately equal to 5R/2, but s is a function of temperature. 
The integral on the right-hand side is quite definite, and is 
equal to the increase in energy-content of the liquid * 
between absolute zero and 0, this increase being the actual 
energy-content if we consider the energy-content to be nil 
at absolute zero. But in any case it is a definite function 
ctf 0, say E (9), and it is implied that E (o) = o. Thus 

L = L +s/~E(0) . . . (21.2.7) 

* If the condensed system at zero is really a solid, no ambiguity arises. 
The energy-content then contains a constant term (the latent heat of 
fusion) in addition to the integral. 



218 STATISTICAL MECHANICS FOR STUDENTS 

Turning to (21 . 2 . 6) we find 
dlogp 




(21.2.8) 

where y is a second constant of integration ; for the moment 
we pass over the point as to the possibility of an indefinite 
quantity arising at the lower limit of the integral since 
E (0)/0 2 is apparently an indeterminate quantity when 
6 = 0. It follows that 

s p log 9 - R logp = ^ + ^ dx - Ry 

U i X 

Hence by (21.2.2) 



which by (21.2.7) 




We can now derive the value of the entropy of the con- 
densed substance at the absolute zero of temperature. It is 

Lt E (9) 

e=o-~r +K ~ s p-^ 

By definition the first term on the right-hand side is s 
where s is the specific heat of the condensed system at 
absolute zero * ; thus 

< (o) = K - 8 p - Ry -f 8 . 

We are now in a position to see the bearing of Nernst'tf 
theorem on this analysis. Cases arise in which several 
allo tropic forms of the condensed solid or liquid exist. Each 
form will have its own individual values of energy-content, 
latent heat, etc., at a given temperature 6. Hence since p, 

* Of course, although E(0)/0approaches a definite value as 6 approaches 
zero, it does not follow that E(0)/0 2 does so ; hence the possibility of 

ambiguity in the integral f ~&(x)jx z .dx. 



ENTROPY CONSTANT OF A GAS 219 

the pressure of the vapour, depends only on 6, it might 
readily be inferred from (21.2.8) that the vapour-pressure 
constant y would depend on the particular allotropic form. 
But any chemical reaction in which one form is converted 
into another would, according to Nernst's theorem, change 
the entropy by an amount which tends to zero as the 
temperature at which the reaction is carried out is reduced 
in value. This means that <f> (o) is the same for all allotropic 
forms. Now experiments on the part of Nernst and his 
collaborators were gradually convincing them that in point 
of fact the specific heats of simple solids and liquids tend to 
zero as 6 approaches zero. Accepting this as an experi- 
mental fact, it appeared that the common value for the 
entropy at zero of any of the various condensed forms is 

K - s p - Ry . . . .(21.2. 10). 

Experiment was moreover indicating more than the simple 
fact that s or d E (8)/d9 tends to zero as approaches 
zero ; it was suggesting that the approach towards the limit 
is so rapid that E (9)/8 2 also tends towards zero, thus making 
the integral in (21 . 2 . 8) a perfectly definite quantity without 
any indeterminateness arising at the lower limit. This is, 
of course, quite in keeping with the theoretical discussions 
of Einstein, Debye, etc., which, however, were historically 
later. That being so, y has a definite value for each form of 
the condensed material. But by (21.2. 10) K Ry must be 
independent of the particular form, and since AC, whatever 
value it receives, is only dependent on the vapour, the con- 
stant y is therefore the same for all the allotropic forms of 
the condensed material. 

It was at this point that Planck stepped into the dis- 
cussion. Thermodynamics, even with Nernst's theorem, 
was still unable to assign any value to (21 . 2. 10) , the value 
of the entropy at zero. The constant y can, no doubt, be 
experimentally determined by applying (21 . 2 . 8) to vapour- 
pressure measurements, but, of course, the entropy -con- 
stant of the vapour, K, is, on thermodynamic grounds alone, 
entirely undetermined. By applying the quantum hypo- 
thesis to the statistical-mechanical considerations of Chapter 



220 STATISTICAL MECHANICS FOR STUDENTS 

VII., Planck pointed out that there was considerable support 
for the view that the value of the entropy at the zero of 
temperature tends to zero as a limit. If this were so, this 
meant that one must assign a definite value to the entropy- 
constant, K, of a vapour or gas, viz., Ry + s p9 where y is the 
experimental vapour-pressure constant, and s p the ideal 
specific heat at constant pressure. It remains to show how 
Planck was guided to this conclusion. His arguments were 
not regarded as quite convincing at first, and an extremely 
interesting discussion on the statistical-mechanical side of 
this matter went on for several years. We shall endeavour 
to summarise these investigations in the following chapter. 

NOTE. For the sake of the readers who are familiar with 
the System of Physical Chemistry, by W. C. McC. Lewis, it 
may be as well to point out that the y of the text above 
is connected with the " characteristic constant " of the 
vapour, denoted in that work by the letter i, by the simple 
relation 

y = i + log R. 
For since 

p = R C 9, 
where C is the concentration, it follows that 



which is the expression on page 76 of Vol. II. (2nd edition). 
The logarithms are, of course, to the Napierian base, e. 
The usual numerical values of the Nernst constant are 
determined on the understanding that 10 is the logarithmic 
base. In that case we must write the value for the logarithm 
of the vapour-pressure as follows 

i L o , s p i a 1 f * E (x) , 

-^ 



, y i + log R 

where c = - = ' - . 

2-3023 2-3023 



CHAPTER XXII 

THE ENTROPY CONSTANT OF A MONATOMEC GAS AND 
STATISTICAL MECHANICS 

22.1 The Magnitude of the Phase-Cell. The reader will 
recall certain remarks made at the beginning of Chapter 
XVI., in which was mentioned the question of adjusting 
probability calculations for a system whose particles may 
pass from a condition involving quantum paths to one in 
which the paths are not so restricted. No answer to such a 
problem can be evolved from purely mathematical a priori 
considerations ; it is necessary to introduce a definite 
postulate and test its results by experiment. It was stated 
that one plausible hypothesis assigns even for a non- 
quantised state a fundamental size to the phase-cell depending 
on a suitable power of h ; in the case of particles with the 
usual three degrees of freedom for translatory motion it will 
be the third power. Now such a postulate naturally leads 
one to speculate if a similar specification of the size of the 
phase -cell may not prove convenient even when alternations 
between quantum and non-quantum conditions are not in 
question. In point of fact, very early in quantum history 
Planck introduced such considerations into the statistical 
treatment of a moiiatomic gas. We hinted as much in 
Chapter VII., where a symbol g was used pro tern, as indi- 
cating some definite magnitude having the dimensions of 
action cubed. There is certainly none of the obvious 
support for this view, which is supplied to the analogous 
hypothesis for internal vibratory motions where spectroscopy 
has provided such powerful aid. Nevertheless, discussion of 
this suggestion of Planck's proceeded apace, and, beginning 
with two rather famous papers by 0. Sackur and H. Tetrode, 
a considerable volume of literature poured out dealing 
with this matter, gradually converging to the view that g 

221 



222 STATISTICAL MECHANICS FOR STUDENTS 

not only exists as a definite magnitude, but that its value 
is A 3 . Indeed, in some quarters attempts have been made to 
justify the postulate that even for the translatory motion of 
molecules in a gas system, quantisation of paths exists, and 
that experimental facts to support such a claim might be 
found at sufficiently high concentrations or at very low 
temperatures. In the former case the space available for 
molecular motion becomes restricted and molecular motion 
would have a zig-zag character which might be regarded as 
a vibration whose central point is gradually shifting. In the 
latter case the average molecular energy becomes very small. 
No very definite experimental results to prove such ' ' degene- 
ration of gases " are available, but it must be admitted that 
the conditions under which quantisation would become 
apparent would be very extreme, and it cannot be said that 
the claim is absolutely disproved as yet. The following 
rather artificial analysis shows how one might formally 
express this view. Regard the gas as enclosed in a cubical 
box of side 1. Disregard intermolecular collisions and con- 
sider all collisions as between molecules and the sides. This 
is a kind of vibration on the part of each molecule with three 
degrees of freedom the amplitude for each degree being L 
The action-integral for the component motion parallel to 
one edge (say the axis of x) is 



\mv x 

J n 



dx, 



which is equal to 2 m v x /, and similarly for the other com- 
ponents. Thus, corresponding to three integers r 1? r 2 , r 3 , 
we would have a particular velocity of translation for a 
molecule given by the conditions 

2 m v x I r 1 h 
2 m v y I r 2 h 
2 m v z I = r 3 hy 

so that the energy of a molecule would have one of a series 
of discrete values such as 

w . 2 , 2 , 2) 

8 m l z ( x 2 3 ; 



CONSTANT OF A MONATOMIC GAS 223 

each value corresponding to a special choice of the integers 

r i> T 2> r 3 

Quite apart from these rather speculative considerations, 

indirect experimental evidence is available for choosing a 
definite size for a phase-cell even in the case of gases and 
putting it equal to A 3 . It exists in connection with the pro- 
blem discussed in the last chapter. 

Reverting to Chapter VII., it was shown that if the number 
of representative points in the respective phase-cells in the 
most probable state are denoted by v l9 v 2 , v 3 , ...... v c , then 

, f 5 , fl , , , (2 TT m k)*k , 3) 

- S v r log v r - n \- log e - log p + log -i - i + _L 

r = l I* 9 A) 

(See equations (7.1.3) and (7.2.1)). 

If W m stands for the complexion-number of the most 

probable state, this means that if we adopt the suggestion 

forgr 

(5 
* (log W m n log n) = R - log 9 - log p 

. . . (22.1.1). 



The reader will recall that in discussing the statistical 
basis of the second law of thermodynamics in Chapters VI. 
and XL. we selected the expression on the left-hand side of 
(22 . 1 . 1) as the entropy of a system. Nevertheless from the 
classical point of view it is clear that k log W m any con- 
stant would serve just as well. However, a reference to the 
conclusion of the last chapter will also remind him of Planck's 
suggestion that the entropy of a condensed system should 
be so chosen as to vanish at absolute zero. One necessary 
conclusion of this would be that the absolute entropy of any 
substance in this sense would be simply proportional to its 
amount. Thus the entropy of a quantity of gas would be 
proportional to n, and this consideration seems at once to 
justify the removal of k n log n from k log W m , as the 
expression on the right-hand side of (22 . 1 . 1) is certainly pro- 
portional to n. But a little reflection raises doubt once more 
since the further addition or subtraction of a multiple of n 



224 STATISTICAL MECHANICS FOR STUDENTS 

(which is a constant) would still leave the expression for the 
entropy proportional to n. On reflecting still more, we are 
led to consider how the removal of k n log n arises in the 
analysis. It takes place when we write for W m the expression 



n n 



and then divide W m by n n , so as to clear n out of the 
expression for " probability " and to obtain a logarithm of 
probability which is proportional to n. But the expression 
just written for W m is an approximation ; why not clear n 
out of the exact expression for W m , viz., 

n\ 



If we do so, we would choose for the entropy not k log (WJn n ) 
but k log (WJn !), i.e., 

k (log W m n log n + n). 

Now this line of reasoning is manifestly weak ; neverthe- 
less it leads to a result which can be tested by experiment. 
The entropy now becomes 

^e-lag P +1 0g VW k + *l . (22.1.2). 

/I 2^ I 

If this be so, then we have at once ascertained a value for, 
the entropy constant of the last chapter, viz., 

-P , (2 77 m k)l k , 5R 
* = Rlog<~ A , - + a 

and from the conclusion there established that the vanishing 
of the entropy of the condensed system at absolute zero 
makes K equal to Ry + s p9 we arrive at the equality 

r = log (2,.U _ _ (22 , 3) 

since s p = 5R/2. 

But y is a vapour-constant which can be determined by 
experiment. In (21.2.8) 



CONSTANT OF A MONATOMIC GAS 225 

is an integral which can be calculated from specific heat 
measurements by the method of quadratures, or by the 
expansion of E (6) as series of powers of 9 whose coefficients 
can be obtained from the experimental data. Thus (21.2.8) 
can be used in conjunction with specific heat and vapour- 
pressure measurements to find y. Equation (22 .1.3) can 
be written 

, (2 77 M) 3 R* 

y = tog -^-' 

where M is the molecular weight of the gas, n the number of 
molecules in a gram-molecule of gas (6-06 x 10 23 ) and R is 
the gram-molecular gas constant (8-32 x 10 7 ). Thus 

y=y + l-5logM . . . . (22.1.4) 
where 

3 5 

y = ' log 2 77 + log R 4 log Ti 3 log h 

2i 2i 

= 10-17. 

If we adopt the base 10 for the logarithms as is general in 
the actual calculation of Ncrnst's constants, we find that 

10-17 , , r , , T 

c = -f 1-5 Iog 10 M 

2-3026 tol 

= 4-41 + 1-5 Iog 10 M . . . (22 . 1 . 5). 

The experimental determination of c in the cases of 
hydrogen, mercury and argon give values approximating to 
4-3, 4-4, 4-45, respectively for the constant term in (22 .1.5). 
The reader will recall that at sufficiently low temperatures, 
hydrogen behaves as a monatomic gas. 

Despite this a posteriori justification by experiment no one 
can feel that the initial argument for k (log W m n log 
n + n) as the entropy in preference to many other expres- 
sions apparently as plausible is very satisfactory. It is 
round this point, rather than the choice of h 3 as the magni- 
tude of the phase-cell, that the discussion has been most keen. 
Planck himself has been most persistent in endeavouring to 
justify the division of W m by n \ by considerations involving 
no other state than that of gas. His latest views are given 
at some length in the fifth edition of his famous book Theorie 



226 STATISTICAL MECHANICS FOR STUDENTS 

der Wdrmestrahlung. Briefly they amount to a statement 
that if we " quantise " the movements of the molecules, it 
does not matter to which molecule we attach a particular 
set of values for the quantum numbers r l9 r 2 , r 3 , and as there 
are n \ permutations of n points possible, division by n ! 
simply embodies this indifference as to the individuality of 
the molecule. Nevertheless many investigators regard such 
arguments as fundamentally weak and hold that only by 
considering some process in which the number of molecules 
in the gas can be varied, can we settle without ambiguity 
the dependence of the entropy expression on n. Such a 
process can be found in the sublimation of a gas from a 
solid state ; we can regard the molecules in the solid as in 
quantum states without postulating quantum states for the 
gas. The following is a short sketch of a line of argument of 
this type due to 0. Stern. 

22 . 2 Stern's Treatment of the Entropy-Constant Problem. 
Stern's paper is lengthy and rather involved. It falls 
naturally into two parts. In one of these, the vapour 
equation (21 . 2 . 8) derived in the last chapter is adapted to 
suit the view that the internal energy E (6) of the condensed 
part of the system is determined by quantum considerations ; 
i.e., thermodynainical reasoning goes into co-operation with 
the hypothesis that quantum states exist in the condensed 
substance (which is regarded as a cubical lattice such as we 
dealt with in Chapter XJ X.), but no hypothesis that quantum 
states exist in the vapour is introduced. The result of this 
train of reasoning is the equation 



, 

r=* 1 2 fj //jiA 

logp = --- _--_ + L log + Slog (jgj + y(22.2. 1) 

where v v ^ 2 , ...... , v^ are the 3/ natural periods of a 

lattice of / atoms as given by (19 .2.7) and (19 .2.8) and v 
is their geometric mean in other words 

" 3/ - Vi . v 2 ...... v 3/ . . . . (22 . 2 . 2) 

R is written for fk. 

The other part of the paper is the derivation by statistical- 



CONSTANT OF A MONATOMIC GAS 227 

mechanical methods of a vapour-pressure formula for a 
cubical lattice and its vapour in statistical equilibrium. In 
this part quantum considerations are not involved, since the 
temperature is regarded as high enough to permit of the use 
of classical statistics. In this part Stern finds that 

. (22 . 2 . 3) 



In this, W =fw, where w is the work required to bring one 
atom from complete rest in the solid into the vapour. No 
constant of integration appears in this equation ; the mole- 
cular model is definite, and so gives the absolute value of the 
vapour pressure. Thermodynamics, on the other hand, only 
gives dp/d9 or d log pjd9 ; it is for this reason that a constant 
of integration appears in (21.2.8) and (22 .2.1). A com- 
parison of (22 . 2 . 1) and (22 . 2 . 3) now yields 

W^L + l 7 ^r 

r-l * 

and the result which we are anxious to justify, viz. 

, (2 77 ra k)% k 
^ ] g A* 

We shall briefly outline the development of each part of 
the paper. 

I. The Thermodynamical Part. In (21 . 2 . 8) let us write 
for a lattice of / atoms 



where, as usual, //, = 

Our task is now to integrate 

% ^ : dx 

r =ix 2 {exp (hv r /kx) 1} 

from o to 0. Considering one term, we have 

dx 



exp (hv r /kx) 1 
dy 



228 STATISTICAL MECHANICS FOR STUDENTS 
where y = hv r /kx and 17 = hv r fk6. This is equal to 



f f e y 

*l \-T \~ 
J, [e y 1 

oo 



At the upper limit, y = GO , log (e y 1), is practically 
log e y or y, and so the integral is equal to 

kr) k log (& 1 ) 

7. 



Now 



log (e^ - 1) - log ^ hv + f /^ 2 ^ ^ 2 

log /it h v + log / 1 + - p. h v + ...... j . 

If the temperature be sufficiently high, p, h v is small 
enough to allow this to be written 

log (e hv 1) log p, h v + - p h v. 

2i 

Hence it follows that 



and thus equation (21.2.8) becomes 

T f 7 If 7 27 ^i- 

L 5 ^ v Av r r-1 r 

log p == H log -) 27 log - 

JL\iC7 2i JL\ j.,-,! A/v 

^ ilv 

L + 27 -r- 



using (22 .2.2) and remembering that R = /fc. This estab- 
lishes equation (22 . 2 . 1). 



CONSTANT OF A MONATOMIC GAS 229 

//. The Statistical-Mechanical Part. The second part of 
Stern's paper will probably cause more trouble to follow, 
since Stern uses the statistical method associated with the 
name of Gibbs. This method will be explained in Chapter 
xxiv., but even at this stage it should not be beyond the 
power of the reader to grasp its essential idea. Indeed we 
have already come very near to its use in Chapters xvin. 
and xix. Recall the fact that we regarded a lattice-system 
of / atoms as a huge molecule and considered the statistics 
of a large number, n, of such systems in an enormous gaseous 
envelope at a given temperature a " temperature-bath " 
in fact. Thus we were dealing essentially with a " Gibbs 
ensemble " (or assembly) of n molecular systems. It is this 
idea that we shall exploit still further here. 

The system is made up of A atoms in an enclosure. Of 
these a atoms will be in a gaseous state, and / will be in a 
solid lattice formation, so that A = a + /, but, of course, a 
and / can vary individually. This system is to be regarded 
as the " huge molecule." An ensemble of these systems, 
n in number, is supposed to be in the temperature-bath at 
temperature d. Thus the systems will exchange energy and 
each system will have a history or travel along a " path " 
in a phase-diagram which will involve the representation of 
3A co-ordinates and 3 A momenta. The system co-ordinates 

will be x v y v z l9 , x a , y a , z a , q l9 &, ? 3 , q sf 

where X T , y r , z r , are the Cartesian co-ordinates of the r th 

molecule in the gas, and q^q^ , # 3 / are 3/ " normal " 

co-ordinates of the lattice, as explained in Chapter xix. 

The momenta of the system are 1? ??, x , , f a , 7? a , a , 

jPi, Pz, p*f where & = mx n rj r = my r , r = mz , and 

p r is the differential coefficient with respect to q r of the 
kinetic energy of the lattice which was shown to be a quad- 
ratic function of the q r involving only squared terms. A 
phase of a system is defined by the assignment of definite 

values to x v # 3/ , g l9 , p 3f . An extension-in- 

phase of the system is defined by a statement that the phase 

of the system lies between x l9 , q 3f , g v 

and x^ + 8x, q 3f + Sq 3f , ( + 8^, p 3f 

Now the probability that a particular system out of the 



230 STATISTICAL MECHANICS FOR STUDENTS 

great number n of these systems is in an elementary phase- 
extension, so defined, is 

I exp (- p, E) So?! ...... 8z a , 8g x ...... S# 3/ 

Sfi ...... S>! ...... 8^ 3/ . (22.2.4) 

where I is some constant, and K is the energy of the system 
in the central phase of this extension which is practically 



...... + , ( } 

2m ' ' 

where (/>(#, >) stands for the quadratic function which is 
equal to the sum of the kinetic and potential energies of the 
lattice, and w is the work required to bring a molecule at 
rest from the lattice into the gas. 

Having settled these preliminaries, let us work out the 
probability that the system is in such a condition that a 
particular atoms are in the gas, and the remaining / par- 
ticular atoms are in the lattice. It will be the sum of the 
probabilities (22 . 2 . 4) for all possible phases, i.e., 

if ...... I" ocp (- /x B) dx l ...... dpv . . (22 . 2 . 5) 

over all phases. 

Our immediate object must be to perform this integration. 
In the first place (22 .2.5) splits into the product of two 
multiple integrals 



J 



dxi ...... dz a d& ..... d a . (22.2.6) 

and 



j ...... J exp [ ~ fji <f>(q, p)} dq l ...... dp 3f . (22.2.1) 

Since [*exp \ - p 2 /2m ! d = ( ) [ V*" dx 

j jj \ p, / j ~ 



it follows that the expression (22 . 2 . 6) is equal to 

e)^ . . . (22.2.8) 



CONSTANT OF A MONATOMIC GAS 231 

where V is the volume of the gas, which can practically be 
considered the volume of the enclosure. 

To work out (22 .2.7) does not require so much " grind " 
as one might imagine, if one refers back to certain expres- 
sions in Chapter xrx. If the reader will look at (19.1. 4), 
(19.2.5), (19.2.6), (19.2.7), (19.2.8), and bear in mind the 
substance of section (19 . 3), he can, if he takes a little trouble, 
convince himself that the energy (f>(q, p) of the lattice in a 
particular phase can be expressed in the form 

2 ' 

where co l5 , o> 3 y are the 3/ natural pulsances of the 

lattice and we have supposed that the co-ordinates q-f 

q 3 j have been changed by a constant multiplier to such 
values that the coefficient of p^ in the kinetic energy is half 
of unity. This makes the calculation of (22 . 2 . 7) fairly 
simple, for since 

f x /I \ /2\ f 

exp I ~LL p 2 } dp { - ) e~ x * dx (2 TT k 6)* 
L,/ V 2* 1 ) \p/ J-oo V ' 

s* rj / i \ / o /. /} \ ' 

i / - 1 oo\7 ( ^ 77 A, 17 ) 

and exp ( ~ ^ ^ q ) dq * 

J _ ^ \ 2 / 60 

it follows that the expression (22 .2.7) is equal to 
f~(277/<;6 



^ rr r i/ x . v 2 i/ 3/ 

(/" f)\ 3f 
-f) (22.2.9) 

v being the geometric mean of the 3/ frequencies v l9 v 2 , 

Combining (22 . 2 . 8) and (22 . 2 . 9), it would seem that the 
probability that the A molecules in the enclosure are so 

* The integration is not actually from q = <x> to q = -+- <x> ; but we 
have on several occasions used the same procedure, since the contribution 
to the integral of the range of q beyond the actual extremes is negligible. 



232 STATISTICAL MECHANICS FOR STUDENTS 

situated that a particular molecules are in the vapour and 
/ particular in the solid is 

But in this conclusion we have overlooked one point. 
We have implicitly assumed that there are not only / par- 
ticular molecules in the lattice, but that they are disposed 
in a particular way, each molecule oscillating about a par- 
ticular position of equilibrium. With these / molecules we 
can make / ! different arrangements out of any one which 
we have just considered, by permuting the molecules 
precisely as if they were / balls in / " pigeon-holes." It 
follows that the complexions are / ! more numerous than we 
thought, and the probability for a particular molecules in 
the gas and / particular in the solid is 



?? /If fl\ z f 

~ I J . .(22.2.10). 



The last step in the probability calculation involves the 
removal of the epithet, " particular. " There are A I/a !/! 
ways of choosing a molecules to go in the gas and / in the 
solid ; so the probability that there are a molecules (any 
molecules) in the gas and / in the solid is 

At * a /1fti\ 

Ll e -i* V a (2irmkO)~*( )*f . (22.2.11). 

a ! \ v / 

This is a function of a and / (really of a only, since a + / 
is constant) and our business is now to find when it is a 
maximum, as this will give us the most probable distribution 
of the molecules between gas and solid, i.e., the distribution 
which we will meet in experiment. To find this, we put the 
differential coefficient of the expression (22 . 2 . 11) with 
respect to a equal to zero or rather more easily, the 
differential coefficient of the logarithm of it. Thus we find 

t o 

-- log a ! p, w + log V + log (2 TT m k 0) 
da 2 

3 log k + 3 log v = 



CONSTANT OF A MONATOMIC GAS 233 

(remember that df/da = 1). Putting log a \ equal to 
a log a a, we find 

i , i TT w , 3 ! /27T m\ , , n 

-loga+logV- +-log(_) +31ogv = 0. 

But the pressure in the gas is given by 



so that 

log p = log a log V + log (Ic 9) 

= - p + | log (27T m) - I log ifl + 3 log v, 
which is just (22 . 2 . 3). 



CHAPTER XXIII 

ENSEMBLES OF SYSTEMS. I 

23 . 1 The Probability Postulate and Dynamics. Through- 
out all the reasoning so far has run the implicit assumption 
that the probability of the occurrence of a particular state 
of a system is proportional to the number of complexions 
embraced in that system. It now behoves us to turn our 
attention to any justification which can be found for this 
postulate. As we pointed out in an earlier passage, the 
complexions of a system can hardly be said to follow one 
another in the same manner as the complexions of a group 
of coins or dice follow one another in the chance of the cast. 
In short, we have to see what dynamical principles have to 
say in this matter. When it comes down to " brass tacks," 
the reader may as well realise at once that we are on very 
doubtful ground indeed if we insist on rigorous logic unless 
we abandon the attempt to be definite about one system 
only and enlarge our field of view to embrace the lives of 
many systems each having the same dynamical character, 
but each having its own individual history which is distinct 
from those of its fellows. 

Paradoxically enough, if we attempt a similar enlargement 
in the case of coin-tossing or dice-throwing, we are entirely 
in the dark experimentally. Many people have tossed 
pennies and found that out of a long succession of attempts, 
practically half are heads and half tails ; but the author is 
not aware of any occurrence in which several thousand 
people threw pennies in the air simultaneously and observed 
the distribution of heads and tails at each throw. Indeed, 
we should have to imagine such a crowd of people each 
tossing several dozen pennies and observing how far the 
proportionality of probability to complexion-number is 
obeyed when the complexion of each system of pennies is 

234 



ENSEMBLES OF SYSTEMS. I 235 

recorded and the facts for the " ensemble " or assembly of 
the systems compiled. Yet if instead of dealing with one 
dynamical system we consider an ensemble of systems, we 
are on very firm ground experimentally and we proceed to 
show why. 

We shall begin with a very simple system indeed our old 
friend, the uniformly accelerated falling particle ; it is a 
system with one degree of freedom ; i.e., its position is given 
by one co-ordinate, viz. , the distance below some definite level. 
But we want to consider an ensemble of these systems ; so 
we shall visualise a shower of rain with the resisting air 
conveniently removed and arrange that each drop has its 
own line of fall with 110 drop exactly in a vertical line with 
another. Under these circumstances many drops may at 
one moment be at the same level, i.e., have the same value 
of the co-ordinate, which we shall denote by q, but we must 
conceive that they have not all the same value of velocity 
or q (q is written for dq/dt) ; in other words, these drops have 
fallen from different heights at such instants as to place 
them now (but not before or after) at the same level. Look 
at this another way ; we know from dynamical theory that 

q=a + bt+~gt 2 . . . . (23.1.1) 

In this equation a and 6 are constants of integration ; their 
values are arbitrary ; they vary from particle to particle 
and give an individuality to each member of the ensemble. 
But g is common to all members. It is determined by the 
field of gravity i.e., the external bodies acting on each 
system affects them all alike. We call it a " parameter " of 
the system. It is obviously possible to choose many sets of 
values of a and b which give the same value of a + bt for a 
given value of t ; the particles involved will all have the 
same value of q at the time t. But since 

q =b+gt . . . . (23 . 1 . 2) 

and since 6 is not the same for this group of particles, their 
velocities are different at this level. In this way we arrange 
that no two particles have the same phase at any given 
moment, phase depending on both position and velocity. 



236 STATISTICAL MECHANICS FOR STUDENTS 

Suppose then we represent this state of affairs on a phase- 
diagram. We shall draw as usual two rectangular axes, 
OX, OY, and represent the phase of each system at a definite 
instant by a point P, the x co-ordinate being equal to q, the 
y co-ordinate being equal to the momentum, mq. As the 
system moves, i.e., as the drop falls, the phase-point P 
moves along a stretch of a parabola in the phase-diagram ; 
this is the " path " or " trajectory " of the representative 
point. This point has a velocity in the phase-diagram, 
whose components parallel to OX and OY are q and p 
respectively. For the moment denote these by u and v, so 
that u q, v p. 

Now it is not difficult to express u and v as functions of x and 
y. Thus the energy of any system of the ensemble is equal to 

mq* + mg (q q), 

where q is a constant (the same for all systems of the 
ensemble). This can be written 

^ i ^P^ + mg(q -q) . . (23.1.3). 

Represent this function by E(g, p). It is easily seen from 
(23 . 1 . 2) that 



3<Z ' 
or u ="-^yl .... (23.1.4) 



V = 



Bx 

D , . Su d 2 E(x,y) 
But since = i-^L' 

ox ox oy 

and 



dy dy dx 

it follows that 



ENSEMBLES OF SYSTEMS. I 237 

This simple relation embodies a result of supreme import- 
ance in our statistical considerations. Since we are con- 
sidering an ensemble of systems, the phase-diagram will be 
covered with a crowd of points or dots, each one moving 
along one member of a family of parabolas. We can conceive 
the drops of rain to be so numerous and so many phases to 
be represented that the representative points are at any 
moment closely packed together in the phase-diagram, 
forming what is practically a dense " cloud " of points 
similar to a two-dimensional continuous fluid, and, of course, 
as the drops fall the cloud moves in the phase-diagram. 
Imagine that any moment the cloud fills a portion of the 
phase-diagram within some closed geometrical curve. Later 
its position has changed ; it occupies a different region 
within another curve, most likely of different shape, but the 
second region has the same area as the first ; that is the vital 
point. As each system of the ensemble passes through its 
successive positions and momenta, the cloud of representative 
points moves about on the phase-diagram, always occupy- 
ing the same extension-in-phase. To prove this, all we require 
is equation (23 .1.5). This, and, of course, a corresponding 
wider theorem, for systems with many degrees of freedom, 
is the great contribution of theoretical dynamics to statis- 
tical mechanics. Its implications we shall elicit presently. 
The proof for the simple one -degree system is not par- 
ticularly troublesome. 

Thus conceive an elementary rectangle of area 8x Sy in 
the phase-diagram whose centre is at the point (#, y), and 
let us calculate the net number of representative points of 
the cloud which move into the rectangle in time St, which 
elapses after a certain instant denoted by t. At that 
instant there will be a definite distribution of density over 
the phase-diagram for the R-points.* We shall denote it 
by p, where, however, we must observe that this density is 
a function of x, y and t, and should really be symbolised by 

* We must distinguish between points in the phase -diagram in the 
ordinary sense and representative points which are conceived to have a 
kind of substantiality and to move about. To signalise this we shall call 
the latter " R-points." 



238 STATISTICAL MECHANICS FOR STUDENTS 

a functional form, such as/(x, y, t) ; for we are not assuming 
that the R-points are uniformly distributed at any definite 
instant, not even over the area in which we considered them 
to be originally placed ; nor do we assume that the density 
remains unchanged at a definite point of the phase-diagram 
throughout all time. It should also be remembered that 
u and v are functions of x and y also, and should be written 
t(x, y), MX, y)* (See (23 . 1 . 4).) 

Consider that side Sy of the elementary rectangle which is 
nearest to the axis OY. Taking the value of u at its mid- 
point (x o Sx, y), we see that the number of R-points 
which cross this side into the rectangle in time St will at the 
instant t lie in a parallelogram whose side is 8y, and height 
St . <f> (x 7! Sx, y) ; for, of course, the component v of the 
velocity contributes nothing towards transport across a line 
parallel to OY. We can conceive St to be so small that this 
height is small compared to Sy, and so the density of the 
points in it at the instant t can be taken to be/ (x I Sx, y, t). 
Thus the number of R-points entering the rectangle by this 
Sy side in time St, is 

Sy /(* - 2 > y> ^ i (* - i ^ y) - ( 23 l 6 ) 

Similarly the number which leave the rectangle across the 
opposite Sy side in time St is 

Syf(x+^8x,y,t)8t<f>(x + ^8x,y) . (23.1.7) 
In precisely the same manner the number which enter 
and leave the rectangle by the lower and upper 8x sides 
respectively, are 



8xf(x, y + J Sy, t) St t(x,y + \ Sy). ' - 
The mathematician is prone to lifting his eyebrows at this 
point, and, when it comes to mathematical rigor, he is justi- 
fied in expressing doubt about our procedure ; but life is 
short, and, in this book at all events, we cannot enter into 
all the mathematical niceties of proof. The reader can rest 
assured that the procedure can be justified if certain restric- 

* In general dynamical reasoning u and v may also involve t ; but we 
are excluding cases where the " geometrical co-ordinates involve the time." 



ENSEMBLES OP SYSTEMS. I 239 

tions as to continuity and finiteness are imposed on the 
functions involved restrictions which we believe are obeyed 
in these applications. The place where the ice is thin is the 
implicit assumption that, for example, </> (x \ $x, y) will 
practically serve as the expression for the velocity at any 
point on the left-hand side Sy of the rectangle. 

Gathering together results (23 . 1 . 6), (23 . 1 . 7) and (23 .1.8), 
we obtain for the net loss of R-points in time from the 
rectangle 

/^. <~\ 

% ;r- !/ (x, V, #(*, V) j ZxSt + dx~\ f(x, y, I) ^(x, y) \ 8y 8t 

dx ' J cy ' ' 

or briefly 

+ ^SxSyS* . . (23.1.9) 



But after time 8t the density of the R-points at the point 
(x, y} becomes 

f(*,y,t) + ^ t f(x, y, t) St. 

Thus the net gain of R-points in the rectangle is 
gfly St 8 X 8y, 

or ^ 81 8x By . . . . (23 .1.10) 

Ol> 

Since results (23 .1.9) and (23 .1.10) must be consistent, it 
follows that 



= 
dt Sx dy 



Another simple step leads to 



dt ox cy \dx dy 

This is the " equatiqn of continuity," which must be satis- 
fied since raindrops are not being created or destroyed. But 
as u and v satisfy (23 .1.5), we conclude that 

I' + ^-f^o . . . (23.1.12) 

' 



240 STATISTICAL MECHANICS FOR STUDENTS 

This is the result of combining continuity with dynamical 
law. 

After time 8t, the R-point which was at the point (x, y), 
has arrived at the point (x -f- u St, y + v 8t), and the 
density of the R-points around it is 

f(x + u 8t, y + v 8, t + 8t) 
i.e., 



or 



which by (23. 1 . 12) is simply p. Thus as the cloud of R- 
points moves about in the phase-diagram, in a manner 
entirely determined by the dynamical behaviour of each 
member of the ensemble of systems, the density around any 
given R-point (i.e., representative of a given drop of the 
ensemble), remains unchanged. Looked at in another way, 
this means that each element of area containing a definite 
number of R-points does not vary in size, and so this must 
also be true for finite stretches of the phase-diagram. The 
reader is warned that the proof does not concern the density 
around any given point of the phase-diagram ; i.e., it is not 
proved that f(x, y, t + 8t) is the same as f(x, y, t), or that 
dp/dt is zero. Put it in the language of Gibbs, we prove 
that "in an ensemble of mechanical systems identical in 
nature and subject to forces determined by identical laws, 
but distributed in phase in any continuous manner, the 
density -in -phase is constant in time for the varying 
phases of a moving system ; provided that the forces of a 
system are functions of its co-ordinates." In this the 
emphasis is on the phrase " varying phases of a moving 
system " ; the proposition is not necessarily true for the 
density around a constant phase. 

To be sure our proof so far has been limited to systems 
with one degree of freedom ; then we can appeal to our 
powers of visualisation to assist the understanding. That 



ENSEMBLES OF SYSTEMS. I 241 

the proof is general enough to cover any system with one 
degree of freedom (and not merely the falling particle) 
provided the force is a function of the co-ordinate, requires a 
very little adaptation at the beginning, which will be appre- 
ciated presently when we take the case of more complex 
systems. But before passing on it will be instructive to 
consider another simple system, viz., the pendulum, since it 
illustrates very clearly a point which is of prime importance 
in the statistical applications. 

In the ensemble of pendula all members have identical 
lengths, and swing in places with a definite value of g. 
In fact I and g are parameters. But the amplitude is an 
arbitrary constant introduced in integration of the equations 
of motion, and varies from member to member of the 
ensemble. The second arbitrary constant is the " epoch- 
angle " which settles at what instants the string is in a 
definite position, say the vertical. Further and this is 
important the periodic time is not the same for different 
amplitudes. The elementary theory of the pendulum, 
which brings out the value 2?r (l/g)* for this, considers only 
infinitesimal swings. The true result is 2?r (Ifg)* /(a), where 
/(a) is a function of the amplitude a concerning which all 
we need to specify at this moment is that it approaches 
unity as a approaches zero, but increases in value with a 
increasing. The angular co-ordinate, 0, and the angular 
momentum ml 2 will be represented in the phase -diagram. 
The R-point of any system will travel round a closed oval 
curve with as centre these ovals will approximate to 
elliptical form for small axes but the time of making a 
complete circuit increases somewhat as the ovals increase 
in size. Take two points, P, Q, on one of these ovals and 
draw normals at them, cutting another oval outside and 
near it in points P', Q'. The figure P Q Q' P' is quasi- 
rectangular. Conceive a cloud of R-points distributed 
uniformly throughout it. As the corresponding systems 
in the ensemble oscillate in accordance with dynamical law, 
the R-points travel round the appropriate ovals. All those 
on one oval go round in the same time, but those on a given 
oval take a little longer than those on one within it. Thus 



242 STATISTICAL MECHANICS FOR STUDENTS 

after one period for those on PQ, those originally on P'Q' 
have not quite arrived back, and similarly for those between 
PQ and P'Q' ; there is a progressive lag as we go outwards. 
So the shape of " cloud-covered " patch is gradually more 
and more distorted from the initial " quasi-rectangularity " 
into a quasi-parallelogram shape, but the area remains 
unchanged. After some time the inner R-points will have 
overtaken the outer, and so the area is stretched like a 
spiral ribbon round between the inner and outer ovals. 
Still later we can have a very thin ribbon indeed spiralling 
many times round as we go along it from inner to outer 
ends. Actually if the R-points were visualised as black 
dots on white paper, the whole space between the two ovals 
would present the appearance of blackness all over, or rather 
a continuous grey ; and so if P Q Q' P' were an area equal 
to l/n of the ring between the ovals it might appear that the 
R-points had now a uniform density equal to l/n of the 
original ; yet this is quite wrong in reality. Density, 
remember, is regarded as a function of x, y, t ; it is obtained 
by dividing a number of R-points in an element of area by 
the value of that area and proceeding to a limit. If we 
conceive this clement of area to be in motion with the 
central R-point of it, we preserve unvarying density. If we 
conceive the element to be fixed, it is alternately empty 
and full of points. In other words, as time goes on the 
" black ribbon " grows longer and thinner, but the " white 
ribbon " in between does the same and maintains its size 
also. The sizes of the two parts maintain the same ratio, 
but of course the R-points are not so compactly situated 
as at the beginning, giving the impression of a uniformity 
of distribution over the whole annulus between the ovals 
which is an illusion. This reference to such an ensemble 
will help us in certain statistical considerations which will 
arise shortly. 

23 . 2 Systems with Two Degrees of Freedom. As the 
simplest illustration of this type, let us consider a particle 
moving in a plane in a field of force due to a centre of 
attraction or repulsion at an origin, and choose polar co- 
ordinates r and 6. Anyone wit)) some knowledge of the 



ENSEMBLES OF SYSTEMS. I 243 

dynamics of a particle knows that the equations of motion 
are 

m (r - r9 2 ) = F 

m | (M) = 
at 

where F is the force directed from the centre. Now the 
kinetic energy is | m(r 2 + r 2 2 ) and the potential energy 
V(r) where F = dVjdr. The total energy is 



The partial differential coefficients of this with regard to 
r and 6 respectively are mr and wr 2 ; these are the 
" momenta " in the general sense used throughout this 
book they happen to be the ordinary linear momentum of 
the particle resolved along r, and the angular momentum 
round the centre. Call them p r and p d . It easily follows 
that the energy is equal to 

(23.2.1) 



and it requires no great trouble to establish that the equa- 
tions of motion above are equivalent to 

dp r == _ dE(r, fl, p r , p e ) 
dt Sr . . (23 . 2 . 2) 



dt 38 

where E(r, 6, p r , p e ) is written for (23 .2.1). 
It can also be seen without much effort that 

dr _ 8E(r, 0,p r ,p e ) 

dt 8 ^ ... (23.2.3) 

dO 



dt dp d 

* It happens that the function E does not contain explicitly, and so 
9E/30 is zero, but this accident does not invalidate the generality of the 
reasoning. 

R 2 



244 STATISTICAL MECHANICS FOR STUDENTS 

In fact, (23.2.2) and (23.2.3) constitute Hamilton's 
form for the equations of motion. 

Conceive now the possibility of a representation in a four- 
dimensional phase-diagram, with r and 6 represented along 
axes OX 1? OX 2 and p r and p e along axes OY 1? OY 2 , so that 
if we denote the components of velocity of an R-point, 
representing some system, along its path in the diagram by 
u v u 2 , v v v 2 , where u l = r, u 2 = 0, v^ = p r , v 2 = p e we 
have 

9 x 2 , y v y 2 ) 



j, x 29 y l9 y 2 ) 
dx l 

9 x 29 y l9 y 2 ) 



dx 2 
From this it follows that 

which corresponds to (23 .1.5) and leads to similar con- 
clusions ; but more of this presently. 

23 . 3 A Rigid Body. This will briefly illustrate a system 
with six degrees of freedom. Three of these are easily 
disposed of ; they correspond to the three ordinary co- 
ordinates of a point fixed in the body, say its centre of mass. 
The other three are a little troublesome to visualise ; the 
reader will be helped by considering the earth in relation 
to the plane of the ecliptic and the fixed stars. These 
latter give us a " system of reference/' viz. 9 the plane and 
some definite direction in it, say from sun to a definite point 
in one of the constellations of the Zodiac. As regards the 
earth we choose a convenient plane in it, and some definite 
direction in that plane, say the plane of the equator and a 
line from the centre to the point on the equator with longi- 
tude zero. Now the plane of the equator at any moment 



ENSEMBLES OF SYSTEMS. I 245 

intersects the plane of the ecliptic in a line, viz., the " nodal " 
line, and there is a definite angle between the planes, the 
" inclination." This angle constitutes one co-ordinate ; 
in the case of the earth it remains nearly constant in time, 
only experiencing a secular periodic variation the " nuta- 
tion " of the earth's axis. Then the nodal line makes with 
the line of reference to the constellation point (they both 
lie in the plane of the ecliptic) a definite angle, which is a 
second co-ordinate ; in the case of the earth there is a slow 
increase of this angle at the rate of about one revolution in 
27,000 years the " precession " of the earth's axis. Lastly, 
the nodal line makes a definite angle with the line on the 
earth from centre to zero longitude on the equator. In 
the case of the earth this third co-ordinate varies by 360 in 
one sidereal day. This method of choosing the " Eulerian 
angles " can be applied to any rigid body in a given frame 
of reference. Works on dynamics show how the kinetic 
energy can be expressed as a homogeneous quadratic 
function of i, y, x, 9, </>, *jj, where 6, <f>, ifj are these 
angles. One important feature is that the coefficients of the 
individual terms may involve functions of the angles, such 
as cosQ, sinO, sin<f>, etc. ; but we noticed that in the previous 
example a coefficient of a squared velocity term might 
involve a function of the co-ordinates (e.g., mr 2 d 2 ). Another 
feature is that product terms in the velocities as well as 
squared terms make their appearance. 



CHAPTER XXIV 

ENSEMBLES OF SYSTEMS. II 

24 . 1 Hamilton's Equations. In considering more com- 
plex systems than in the previous chapter, we can realise 
that the complete configuration of a system will be given 
by a suitable number of co-ordinates. Three of these would 
specify the position in a reference frame of some definite point 
of the system, say its centre of mass. For each particle 
in the system not rigidly bound to other particles we would 
have three further co-ordinates in a frame of reference 
having this point as origin. If any rigid bodies entered 
into the constitution of the system, each of these would 
introduce six co-ordinates. These co-ordinates (in most 
practical applications they are Cartesian or polar co- 
ordinates, with Eulerian angles in the case of rigid bodies) are 
denoted usually by symbols such as q l9 q 2 , . . . , q^ where j8 
is the number just necessary and sufficient for complete 
description of the configuration of the system. The Car- 
tesian co-ordinates of every particle or element of volume 
of the system can be expressed as functions of q v # 2 , . . . , 

ft; e '9-> 

XT = <l>r til, ? ft) 

Vr = tr til, V* ft) * 

*r = X r til, ?2> ft)- 

The velocity of each particle or element of volume is given by 

d(f> r . dJ> r . d6 r . 

^ = ~q^ + ^T-<l2 + + 3^, 

dq l dq 2 dq ft * 

and two similar equations. 
Recalling the expression for kinetic energy, viz., 

-Zm r (x* + 2/r 2 + 2> 2 )> it is clear that the kinetic energy of 

2* 

* In certain special cases t is also involved in the functional form, 
i.e., the geometrical equations involve the time explicitly. 

246 



ENSEMBLES OF SYSTEMS. II 247 

the system is given by a homogeneous quadratic function of 
the generalised components of velocity q i} q 2 , . . . qp, the co- 
efficients of the squared and product terms being in general, 
not constants, but functions of the co-ordinates. Denote 
this by 

2 ( a n tfj 2 + + 003 <lf? + 2 12 qi q. 2 + . . .) . (24 . 1 . 1) 
The generalised components of momentum, denoted by 
Pii P& - P/3> are obtained by differentiating this expres- 
sion with respect to q l9 jo, . . . , j^, respectively. Thus 

Pr = r 1 ?1 + r2 </2 + - - - + ^ ft 

where it is implied that a rs = a sr . It is now possible to 
express the kinetic energy in terms of the components of 
momentum. It is also a homogeneous quadratic function 
of these, such as 

\(b llPl * + . . .+b ftft p f * + 2bup l p + . ..) (24.1.2) 



where the coefficients b rg are in general functions of the 
co-ordinates. 

The potential energy of the system, whether arising from 
inter-action between its parts or from action between these 
and external bodies is a function of the co-ordinates V^, 
?2 ft)> or briefly V(^), involving among its parameters 
any quantities necessary for specifying the relations of the 
system to such external bodies. The sum of this function 
and the function expressed in (24 .1.2) is called the 
" Hamiltonian function " of the system. Let us denote it by 
E (?i, ? 2 ?0> Pi P^ Pp), or briefly E (?, p) ; its 
value for given values of the q r and p r is, of course, equal to 
the energy of the system when in that phase of configuration 
and motion. The importance of this function lies in the 
fact that nearly a century ago the Irish mathematician, 
Sir William Rowan Hamilton, demonstrated that the equa- 
tions of motion of the system could be written in the form 

dq, = 9E (q, p) 

dt * P > ... (24.1.3) 

dp, = __8E(g,y) V ; 

dt d 



248 STATISTICAL MECHANICS FOR STUDENTS 

there being /? equations of the first type, and p of the second. 
Even in the case of non-conservative systems where the 
forces cannot all be derived from a potential function, 
provided such forces are functions of the co-ordinates, we 
can write the equations of motion 
dq r _SE(q,p) 

dt **> . (24.1.4) 



where the ft functions Q, (q) of the co-ordinates depend on 
the forces which are not included in the potential function 
V (q). Representing dq r /dt, or q r , and dpjdt, or p, , by u r and 
v r respectively, we see that each of the u r and v r are functions 
of the phase quantities q lt g 2 , . . . q ft9 p l9 p 2) . . . p ft . 
Further, it follows from (24 . 1 . 3) or (24 . 1 . 4) that 



dq r dq, 



dp, dp, Sq, 

Hence we obtain 



. . .(24.1.5). 

dq r dp r 

24 . 2 Liouville's Theorem. We can now prove the general 
theorem mentioned in the last chapter for an ensemble of 
systems all identical in nature but of any degree of com- 
plexity. Each system of the ensemble has its own individual 
history, but the initial conditions of each system have been 
so chosen that at time t, the phase of any system is different 
to that of any other. Yet the number of systems in the 
ensemble is so great that within a finite extension-in-phase 
there are an enormous number of them. An element of 
this extension-in-phase can be specified by the limiting 
values 



ft + S ?2>#i + 8 ^i" "Pf* + 



ENSEMBLES OF SYSTEMS. II 249 

meaning that a phase in this extension is given by q l 
a v . . . q ft dp Pib v . . . p ft bp where a l9 . . . b ft 
are 2 /J small positive quantities less in numerical value than 
i 8j lf . . . ^ Spp respectively. The number of systems of 
the ensemble within this element at time t is 



where p is a function of the q r and p r and Z, say / (q, p, t). 
To find the number at time t -f- 8t we must calculate the net 
gain of systems within these limits in time t, due to the fact 
that the phase of each system has altered in the interval. 
If 8t is chosen small enough, systems which have at time t 
phases given by 

0i- 5 8 ?i - *n 02 2> 9 ft *v Pib v . . . Pfi b ft , 

u 

and which are therefore outside the element at time t will be 
within the element at time t + 8, provided h is not greater 
than (u l | dujdq^ SqJ St. On the other hand those 
whose phases at time t are given by 

?1 +-3 ^ ~ k l> ?2 2 ' ' & ft /3' JPl 6 1 ' Pft V 

and which are within the element at time 2 will be outside it 
at time t + S provided k^ is not greater than (u l + 9^ 1 /3g 1 . 
Sg^) 8^. The net loss of systems to this element in the 
interval S on this account will be given by 

~ Ul 8 ?i & 8g 2 ... 8^ 8p l . . . 8p ft . 

C$1 

Proceeding in this way the complete net loss in the 
interval Stf is given by 

S* 8q l . . . 8^ 8^! ... Sjpj X 



But the net gain is also 
? 8* S . 



250 STATISTICAL MECHANICS FOR STUDENTS 

Hence, since these expressions must be consistent we have 
the fundamental result 



...... _ Q ^ (24.2.1) 

dt dq dp ft 

But by reason of (24 . 1 . 5) this reduces to 

. . (24.2.2) 



. 

ot cq r 

Now the system of the ensemble which was at the phase 
<?!> Pp at the time t is at the phase q l + u 8t, . . . 
Pp + Vp St at time t + 8t and the density-in-phase of the 
ensemble at this phase is f(q + uSt, p + vSt, t + or 



which, on account of the condition (24 .2.2), is just p. 
This establishes the general theorem referred to in the 
previous chapter. Expressed in other words this means 
that if a particular system of the ensemble is at a phase 
?i,... pp at time t and p is the limit of the number of 
systems within an element of cxtensioii-in-phase q l i a l9 
- Pp bp a ^ that time divided by 2^ a l . . . a^ 2& b l . . . 
6^ when a 1? . . . b^ are indefinitely reduced, and if the 
particular system is at phase q' l9 . . . p'^ at time t' and p 
is a like limit at time t', then p = p' . The use of the methods 
of the calculus imply that we must conceive the number of 
systems in a finite extension -in -phase to be increased without 
limit, just as in hydrodynamics we conceive a fluid to be a 
continuous medium. There is still another way of looking 
at this important proposition. We can assume that all 
the phases in a certain finite extension-in-phase satisfy the 
condition that each of them make a function (f> (q> p) of 
?!,... Pp negative or zero in value. Using geometrical 
language we can say that those phases which make the 
function zero are on the " hypersurface," </> (q, p) = 0, and 
the others " within " it. Suppose a number of systems of 
the ensemble are at these phases at time t. At time t' they 
are at other phases, and from the dynamical equations 
?i'> P'ft can be expressed as functions of q ft , . . . 



ENSEMBLES OF SYSTEMS. II 251 

p ft and ' t, so that we can obtain a function i/j (q', p') 
of q' l9 . . . p'p which is equal to < (q p). Then all the 
systems at time t' will be within the hypersurface 

ifj (q, p) - 0, 

or on it. Liouville's theorem then states that the integra- 
tion of dq . . . dq ft dp l . . . dp ft throughout the extension 
determined by </> (q, p) being zero or negative yields an 
integral equal in value to that obtained by integrating 
dq t . . . dq^ dp-^ . . . dp^ throughout the extension deter- 
mined by i/j (q, p) being zero or negative. Thus there is a 
conservation of the extension-in-phase occupied by a group 
of systems distributed continuously in a finite extension to 
begin with. 

One word of caution is necessary. The actual systems 
which lie within the " rectangular " element q l a l9 . . . 
Pp it ft/3 at time t, are not in general the same as those which 
lie within the rectangular element q\ a v . . . p' ft fys 
at time t f . The number is the same, but identity is not 
necessarily preserved. Put in geometrical language, the 
" shape " of <f> (q, p) ~ is not necessarily the same as 
that of (q, p) = 0. If a rigid body moves in our per- 
ceptual space, certain relations must hold between the 
co-ordinates x, y, z of a particle of it at time t, and the 
co-ordinates #', y', z f of the same particle at time t' by reason 
of the fact that the moving thing is rigid and preserves its 
shape. But no analogous relations hold between q l9 . . . p ft 
and q' l9 . . . p'^ ; at all events Liouville's theorem makes 
no such claim ; it simply proves conservation of the exten- 
sion. 

Last of all we have clearly not proved that dp/dt is zero, 
or that the density-in-phase about a particular phase is 
conserved. 

It is, of course, understood in all this that no system of an 
ensemble acts on any other system ; each pursues its pre- 
destined dynamical " path," under the influence of its own 
internal interactions or of the actions of external bodies, 
and it should be carefully noted that the disposition of 
external bodies is supposed to be alike for each system. 



252 STATISTICAL MECHANICS FOR STUDENTS 

Each member of the ensemble is really system plus external 
bodies. At a given instant the phase-quantities which deter- 
mine the configuration and motion of the various parts of 
a system differ from the member to member of the ensemble, 
but those parameters which determine configuration and 
motion of the external bodies are alike for all members. 
These parameters may, of course, change in time, but at a 
given instant they have identical values for all the members. 
As an example of this we have the ensemble of rain-drops or 
pendula treated in the last chapter. Another example, an 
ensemble of gas systems, is treated on the assumption that 
each one is enclosed in one of an enormous group of identical 
vessels, i.e., each one is subject to an external field of force 
which is alike for all. 

24 . 3 Microscopic States and Their Probabilities. A 
" microscopic state " of a system is defined by stating that 
the phase of the system lies within a small rectangular 
element of extension-in-phase about a particular central 
phase. 

Now the successive phases of an actual system must, by 
the laws of dynamics, satisfy the relation 

E (q, p) = constant. 

These are, in fact, in geometrical language, the energy 
hypersurfaces, and the dynamical path lies on one of them. 
It is clear, therefore, that the phases possible to a system 
with a given energy do not occupy an extension with the 
full dimensionality, 2/?, of the ensemble. We can, however, 
conceive an ensemble of systems limited in such a way that 
no system has an energy less than a definite value X, and 
none has a value more than a value X -f 8X, greater by an 
infinitesimal amount. Imagine also that at any instant 
the systems of the ensemble are so distributed in phase 
that their density-in-phase is uniform throughout the 
extension bounded by the hypersurfaces. 

-E(q,p)=X 

E (ff, p) = X + 8X . . . (24 . 3 . 1). 
Liouville's theorem then shows that this state of things is 



ENSEMBLES OF SYSTEMS. II 253 

perpetual. In fact, since at time t, the 9p/3g f and 
are all zero, it follows that 



also. Thus the ensemble is in statistical equilibrium. 
There is no tendency for the systems to crowd into any 
particular part of the extension defined by (24 . 3 . 1) at 
the expense of other parts. 

Imagine that we are endowed with the power of making 
a choice of one system out of the ensemble. We are just 
situated in a position similar to a person asked to select 
any ball out of a number of boxes each containing the same 
number of balls. The chance that the choice will fall 
within a certain box is I/A if there are A boxes. So the 
chance that we select a system within a given microscopic 
state is I/A if the extension defined by (24 . 3 . 1) is divided 
into A equal elementary extensions. Or to put it another 
way, all the different microscopic states of the ensemble 
have equal probabilities. Further, this chance is not 
upset with lapse of time, as would have been the case had 
Liouville's theorem not been true, for then systems would 
have crowded in increased numbers into certain regions of 
the complete extension, giving, therefore, after a time, an 
enhanced probability that one of the microscopic states 
within this region would contain the system chosen, the 
chances elsewhere being on the other hand diminished. 

An actuary, it is well known, when preparing tables for 
insurance or other purposes does not base his calculations 
on the facts concerning one or two individuals, but on 
averages extending over large groups representative of each 
section of a community of individuals. Our point of view 
at this stage is very like that of the actuary. We really 
have not been basing our statistical calculations in the 
preceding chapters on the behaviour of a single molecular 
system, but on the average behaviour of a large representa- 
tive group of systems, members of an ensemble. Each 
system is composed of the same number n of molecules, all 
of the same kind, each with / degrees of freedom. If more 



254 STATISTICAL MECHANICS FOR STUDENTS 

than one type is present, each system contains n molecules 
of one type each witH / degrees of freedom, ri of a second 
type each with /' degrees, and so on. The number of 
degrees of freedom of the system is nf, or nf + n' /' -J-, etc., 
as the case m&y be. Thus in the case of one type of molecule 
^ #ii> #12 -j (?i/ denote the co-ordinates of the first 
molecule in the system q 2l , # 22 , . . ., q 2f those of the second 
molecule, and so on, then the co-ordinates of the system 

are q lv , q nf , the number being nfin all; in short, 

the suffix j8 of the preceding portion of this chapter is nf 
(or nf + n' f + etc.). It is very necessary to be on guard 
here against a fatal misconception. In a sense a molecular 
system is an ensemble, for it contains an enormous number 
of molecules, and each molecule is in itself a dynamical 
system. But that is not the sense in which the word 
ensemble is used here. For one thing, the molecules inter- 
act on one another, and Liouville's theorem certainly does 
not hold for a molecular system regarded as an ensemble of 
molecules. The systems must be regarded as independent 
of each other, each pursuing its own dynamical path un- 
interfered with by any other system if Liouville's theorem is 
to be true. It is as well to introduce here a simple notation 
to indicate the distinction made. We have in the past 
spoken of representing the momenta and co-ordinates of a 
molecule in a phase-diagram or " phase-space " with 
dimensions 2/ ; we shall refer to this as the M-space. But 
now we must also think of a phase -space in which a whole 
molecular system is represented by a single " point." Such 
a space has a dimensionality 2/3, where /? = nf (or nf + 
ft'/' + etc.). We shall call this the G-space. (Gibbs' 
phase-space.) It is in this space that an ensemble of 
systems is represented by a " cloud of points." 

A microscopic state of the system is obtained by assigning 
narrow limits to the /3 co-ordinates and /? momenta of the 
system. A moment's thought will show that this is tanta- 
mount to assigning a complexion to the molecules of a 
molecular system. Thus if all the systems are considered 
to have values of energy practically the same for each, i.e., 
between narrow limits such as X and X -f SX, then each 



ENSEMBLES OF SYSTEMS. II 255 

complexion of the system, which is indicated by assigning 
n l particular molecules to the first phase-cell in the M-space, 
n z to the second, . . ., n c to the c th , has, as it were, one 
particular phase-cell in the G-space within the shell between 
the two hypersurfaces as its home. A statistical state of 
the system is therefore associated with a particular group 
of phase-cells in the G-space, the number in the group 
being n\/(n- L \n 2 \ . . ., n c !). But we have seen that all 
but a relatively insignificant number of complexions are 
embraced within a statistical state in which the number of 
molecules in an M-phase-cell, corresponding to molecular 
energy , is proportional to e~^ e (where 3/2/x is the average 
kinetic energy of translation of a molecule) or agrees very 
closely to this distribution. This means that if we fill the 
shell in the G-space with a uniformly dense cloud of repre- 
sentative points, then on choosing one at random there is an 
enormous probability that we shall select one which repre- 
sents a system in, or very near to the Maxwell-Boltzmann 
distribution of molecular co-ordinates and momenta. 
Moreover, and this is where Liouville's theorem comes in, 
this state of affairs if arranged for initially is not upset in 
course of time. The distribution in the G-space remains 
uniform, and so on, returning to make a choice at a later 
instant the chances are still enormously in favour of select- 
ing a representative point associated with a system in or 
near the Maxwell-Boltzmann distribution. In other words, 
this distribution is a normal property of the system. The 
word " normal " was suggested by Jeans as a convenient 
epithet for any property which is common or nearly common 
to every member of the ensemble distributed in a uniform 
manner throughout a region of the G-space. As giving 
point to the warning uttered above, the reader will observe 
that the distribution of representative points in a uniform 
manner in the G-space does not lead to uniform distribution 
of M-representative points in the M-space, except in rela- 
tively few cases. An enormous preponderance lies with a 
very non-uniform distribution in the latter space. 

This is essentially an actuarial process, and actuaries 
know that, for the purposes involved, although the life of a 



256 STATISTICAL MECHANICS FOR STUDENTS 

given individual may in its course be seriously at variance 
with the average life of the whole community, yet things 
will work out alright in the end on the assumption that it 
is the same in certain particulars. This is essentially the 
hypothesis we introduce at this point of our reasoning. 
Dynamics can carry us no further. It has had its say. 
A given molecular system might behave very differently 
from one which is always in or near the Maxwell-Boltzmann 
distribution. There do exist relatively very small regions 
of extension-in-phase in the G-space where matters are very 
different for any system whose representative point happens 
to be there, and we cannot definitely deny the statement 
that once there it will always remain in such a region, or 
that the greater part of its path may lie in such regions. 
All we can say is that it is unlikely. But the reader must 
carefully note that in carrying over what is undoubtedly a 
proven statement for the average behaviour of all the 
systems in an ensemble at any instant, for application to a 
single system, and asserting that such is the average be- 
haviour of a single system throughout a long time, we are 
taking a step which is undoubtedly plausible, but which 
cannot be definitely proved. The hypothesis was called by 
Maxwell the " principle of continuity of path " and by 
Boltzmann the " crgodic hypothesis/' It amounts to an 
assumption that a given molecular system will in course of 
time pass through all the complexions consistent with its 
energy, or at all events through a large group of them 
sufficiently representative to them all, before returning to 
or very near to some original phase, and thereafter pursuing 
the same path as before. Yet it must be admitted that 
the dynamical systems which have been most thoroughly 
worked out in detail by the mathematicians, viz., the astro- 
nomical systems, give little or no support to this view. 
Indeed we can realise that our own solar system, if absolutely 
free from all external influence and left to the internal 
gravitational action of its parts, would go on cycle after 
cycle, much as it is, never approaching in successive ages 
unlimited phases which we can conceive it to have by tum- 
bling planets and orbits about in our imagination. There is, 



ENSEMBLES OF SYSTEMS. II 257 

of course, an intuitive feeling that the very complexity of 
molecular systems and the multiplicity of their parts favour 
the hypothesis, but this feeling does not constitute 
proof. Perhaps the most helpful idea is contained in the 
undoubted facts that any molecular system can hardly, in 
practice, be said to be free from other influences than those 
ostensibly introduced in theoretical discussion. For example, 
even in a thermostat slight fluctuations of energy must go 
on in any system ; this means that though for a time the 
system may pursue a definite path on one energy hyper- 
surface, presently it will be " displaced " to another on a 
neighbouring hypersurface ; and although any one path 
might be far from traversing a sufficiently representative 
group of microscopic states, yet the fortuitous shifting 
from path to path may produce this result in the long run, 
and effectively prevent any system from remaining in a 
freak region of extension-in-phase for anything but a 
negligible period. 

The " freak " regions are relatively very small, and any 
system might be compared to a man blindfolded and left 
to wander at random in a large field. The chance that he 
would walk into a small circle drawn in a particular spot on 
this field is small. If in the circle, the chance that he would 
remain in it and not wander out of it is also small. There is 
no impossibility in his walking into the circle and " depress- 
ing his entropy," but in all likelihood his entropy will soon 
rise again, and he will walk out of it. 

Let the reader also recall the example given in the previous 
chapter, where an ensemble of simple systems, compactly 
grouped in the beginning, " spread themselves out" with- 
out contradicting the law of density-conservation ; so 
that a choice at the beginning would have given one of a 
rather restricted group of phases, while at a later time a 
choice would be made from a group of phases more repre- 
sentative of all the phases possible. The reader should in 
this connection consult Gibbs' Elementary Principles, 
Chapter XII. 

The ensemble method is very powerful in dealing directly 
with problems where we are concerned with a system which 



258 STATISTICAL MECHANICS FOR STUDENTS 

is not isolated, and it is a question of the probability that it 
contains a certain energy, or rather that its energy is within 
certain narrow limits. This probability is proportional to 
the number of microscopic states consistent with these 
energy limits, i.e., to the magnitude of the extension-in-phase 
bounded by the energy-hypersurfaces corresponding to the 
upper and lower limits. Of course, when quantum assump- 
tions come in we have to modify this strictly classical 
result in a manner which will be obvious on glancing back 
at Chapter xvi. A microscopic state surrounding a 
particular phase is the element in the classical discussion ; 
in the quantum discussion the element is the whole succes- 
sion of phases in a given quantum state. The probability 
in the one case is the magnitude of the G-phase-cell of 
dimensionality 2/3, i.e., having physical dimensions equal to 
action raised to the power /?. In the other it is h ft . To 
illustrate this procedure we shall return to a fresh discussion 
of the entropy-constant problem, but before doing so we 
must point out the importance of choosing momentum in- 
stead of velocity to indicate phase. We have indulged in this 
practice throughout, and the cause is clear. The elegant 
form of Hamilton's equations, leading as they do to the 
fundamental statistical theorem on density -conservation, 
depend entirely on choosing the p r , and not the q r , as 
indicating the motion aspect of the phase. In a phase - 
diagram in which q r and q r are represented there is no 
conservation of density of representative points around a 
moving phase, and no simple basis for probability calcula- 
tion presents itself. 

24 . 4 The Entropy-Constant Once More. Let us consider 
once more a system of n simple monatomic molecules 
represented by a point in a G-space of 6n dimensions. Sup- 
pose the system to be in a state in which a particular atoms 
are in the gaseous state, and the remaining b particular 
atoms are in a solid cubic lattice, and arranged in a particular 
way in the lattice. 

To simplify the discussion we will also assume that the 
lattice is also in its lowest quantum state, so that its energy 
is be where is a constant. In order that any molecule 



ENSEMBLES OF SYSTEMS. II 259 

may leave the lattice and enter the gaseous state its potential 
energy must be increased by an amount w. Hence, if 
K is the kinetic energy of the gaseous molecules, then 
K + aw + 6e must lie between x and X + ^X' Regarding 
for a moment the gas molecules as a separate system with 
3a degrees of freedom, the probability that it will contain 
an amount of energy whose value lies between -^ aw be 
and x aw be + <>x * s proportional to the magnitude of 
the extension in a Ga-dimcnsional phase-space bounded by 
the energy hypersurfaces 

(^ + ^ + i 2 + + a + ^ + C 2 ) 

= 2 m (^ aw be), 

and 



= 2 ra(x aw be 

It will be shown in a note at the end of the chapter that 
this is 

(277-m) c v a { aw _ & y-i g (24.4.1) 

(c - 1) ! V * ' * V ' 

where v is the volume of the gas, practically of the containing 
vessel, and c = 3a/2, a being assumed to be an even number. 
Since we have limited the lattice to one quantum state, 
viz., the lowest consistent with the number of molecules in 
it, the probability that it is in this state is given by an 
extension in a 66 dimensional phase-space. This is h* b 
according to the quantum postulate just referred to. Thus 
the probability that the a particular molecules are in the 
gas and the remaining b in the lattice arranged in a particular 
way and in their lowest quantum state is given by an 
extension in a Pm-dimensional phase-diagram, whose magni- 
tude is 

h*(2*mY( X --b*Y- 1 8v / 24 4 2] 

(c 1) ! A v ' ' '* 

The probability that any a molecules may be in the 
gaseous state, and the remaining 6, still arranged in one 
way in the lowest quantum state, in the solid is obtained by 
multiplying (24 . 4 . 2) by n \/(a ! b !), which is the number of 

82 



260 STATISTICAL MECHANICS FOR STUDENTS 

ways of choosing a molecules out of n. But this has still to 
be multiplied by b \ to introduce all those complexions 
which correspond to any arrangement of the b molecules in 
the lattice. Thus the probability that there are a molecules 
in the gaseous state out of the total number ?i, is 

_ - h v a (2rr m) c (v - aw - be)*" 1 Sy . (24 . 4 . 3). 
a!(c-l)! * * 

The most probable value of a is found by finding a maxi- 
mum value of this expression, subject to the condition that 
^ and 8^ are constant. Referring to the expression 
(24 . 4 . 3), with n ! and 8^ omitted, as P, we, as usual, 
take logarithms and use Stirling's theorem. Thus, 

3# 

log P = 3 (n a) log h -f- a log v + IT 1 (^ m ) 



If we now write down the condition that d log F/da is 
zero we find, after a little re-arrangement, that a is deter- 
mined by the equation 

5 3 3 e w 

-log a - log fv ne 4- a (e iv)\ -- - - ; - --- : 

2 fe 2 6LA - v yj 2 ^~H + a (~ w) 

3 3 3 

= - 3 log h + log v + 2 log 277 m - ^ log-- 

If we now connect this statistically most probable dis- 
tribution with the macroscopic experimental state, for which 
the given total energy ^ determines, with the volume v and 
the number a, the temperature 0, we know that ^ aw 
6e, being the kinetic energy of the gaseous molecules, is 
(3/2) a k ; further we know that the pressure p of the gas 
is a k 6/v, since v is practically the volume of the gas. Hence, 
since log p = log a log v + log k9 we readily find that 

3 , 7 n e w 
Ioga--logk9--^j- 

3 

= 3 log h + log v + - log (2?7 m) 

* |n practice a is so large that we can replace 3a/2 1 by 3a/2, 



ENSEMBLES OF SYSTEMS. II 261 

w e 5 3 

or logp = - + + g log (0) + - log (277 m) - 3 log h 

w e 5 (277 m ft) k 



This is the vapour-pressure equation for the simple model 
we have chosen. It clearly corresponds to the general 
thermodynamic equation (21 . 2 . 8) in Chapter xxi. 

w 1 , L 

-- - corresponds to - 

ku R0 

ice " " 



5 S P 

2 " " R 

, (2rr m k) 3 - k 
7 log fe8 

The method which is due to Ehrenfest and Trkal is some- 
what less laborious than Stern's, and differs from his most 
signally in the fact that it treats the statistical considerations 
on a quantum basis, while Stern's treats the statistics on 
classical lines, while applying quantum methods to the 
thermodynamical investigation. 

24 . 5 Gibbs' Canonical Ensemble. In order that an 
ensemble may form the actual basis for a statistical treat- 
ment of a dynamical system, it is necessary that the ensemble 
should be in statistical equilibrium. This means that the 
density -in-phase of the ensemble round a given phase must 
not change in time, i.e., that dp/dt must be zero. Now this 
condition is not generally true. Liouville's theorem demon- 
strates that the density around a dynamically-moving phase, 
i.e., around the " point " representative of a given system, 
is constant, i.e., that dp/dt + 2 q dp/dq r + 2 p r dp/dp r is 
zero. It is therefore necessary for the statistical equilibrium 
of the ensemble that the condition 



= . . (24 . 5 . 1) 



262 STATISTICAL MECHANICS FOR STUDENTS 

be satisfied. If this is so, any assumption that the prob- 
ability of a given microscopic state being occupied by a 
single system at a chosen instant is proportional to the 
density of ensemble at the central phase of that microscopic 
state, will not be upset with lapse of time, as would be the 
case if the conditions of density around given phases kept 
altering. 

The simplest type of ensemble satisfying (24 .5.1) is, 
as we have seen, one uniformly distributed in phase, but it 
is by no means the only one. If, for instance, the density 
is chosen initially at each phase to be a function of the 
energy associated with that phase, (24 .5.1) is satisfied 
in cases where the energy of the system is conserved ; for if 



p = 

where ^ i g the value at the phase of the Hamiltonian 
function E (g, p) of the system ; then 



dq r d x ' dq r 
and 



dp r d x dp r 

and so 



dp , . 9\ _df( x ) ( dg 

dq r Vr Z'pJ d x (dt~ dq r " dt ~dp r 



. 

d% dt 

= 0.* 

If we wish to maintain other suitable mathematical 
conditions for the ensemble as well as that of statistical 
equilibrium, the function f (\) must be subject to other 
conditions . Thus / (^ ) should not be infinite anywhere . The 
reader may think this condition strange in view of the fact 
that we have several times postulated an increase of the 
number of systems in the ensemble to an enormous value so as 

* Note that this result is only true for conservation of energy. In other 
cases the right-hand side is 2 Q r q r . See Hamilton's equations. 



ENSEMBLES OF SYSTEMS. II 263 

to produce conditions approximating to a " continuous fluid " 
of representative points in the G-phase-space. But this is 
not really the same thing. The function f (%) must n t 
become merely an indeterminate infinity anywhere as it 
would do if it were put equal to, say -%~ l or ^ an X> or so 
forth. Writing it as N ^ (^), where there are N systems in 
the ensemble, we see that 



over the whole phase-diagram must be unity, and this 
clearly debars us from the choice of many functions. Thus 
</> (^) cannot be simply proportional to ^, nor can it have 
the same constant value everywhere, although we may 
give it a suitable constant value between a pair of energy 
hypersurfaces and make it zero elsewhere, as, indeed, we 
did in the preceding considerations. 

Gibbs points out that an ensemble distributed in such a 
way that the density-in-phase is at all phases proportional 
to 

exp f ^ ^M . . . (24 . 5 . 2) 

where is a constant, which he calls the " modulus of 
distribution," " seems to represent the most simple case con- 
ceivable, since it has the property that when the system 
consists of parts with separate energies, the laws of distribu- 
tion-in-phase of the separate parts are of the same nature 
a property which enormously simplifies the discussion, and 
is the foundation of extremely important relations to 
thermodynamics." Such an ensemble he names " canoni- 
cal." An ensemble with a constant density between two 
very near energy-hypersurfaces, and a density which is 
zero elsewhere, he calls " microcanonical." 

A considerable part of Gibbs' " Elementary Principles " 
is devoted to working out the properties of canonical 
ensembles and showing how the rational foundation of 
thermodynamics is related to them. There is no need to 
go into the matter very fully here. Gibbs' work is certainly 
not easy reading for a beginner ; but the reader who has 



264 STATISTICAL MECHANICS FOR STUDENTS 

struggled through the previous pages to this point is no 
longer a beginner. The author feels that he can be safely 
left at this point to pursue his further studies in this classic 
of Statistical-Mechanical theory, and will bring this book 
to a close with a few remarks on the canonical ensemble. 

In the first place one must carefully guard against con- 
fusing the Maxwell-Boltzmann law of distribution of 
molecular co-ordinates and momenta in a system of mole- 
cules with Gibbs' specification of a canonical ensemble of 
dynamical systems. On the assumption that the probability 
of a statistical state of a molecular system is proportional to 
the number of complexions consistent with it one can 
deduce the result that the number of molecules in a given 
element of phase-extension (i.e., M-phase) is proportional to 
e~^ where e is the energy of a molecule at the central phase 
in the most probable state. But this statement implies no 
necessity that this is true all the time. There are un- 
doubtedly fluctuations from this state, such fluctuations 
being less marked the greater the number of molecules in 
the system. Gibbs on the other hand considers ensembles 
of systems. These systems are not necessarily molecules 
nor even groups of molecules, although his work finds its 
most ready application in dealing with systems of molecules. 
His system is any group of objects subject to dynamical 
law. Further, the systems of the ensemble are isolated 
from one another. Whatever law of distribution is laid 
down for them, they follow Liouville's law, and under 
certain conditions this implies statistical equilibrium with 
no fluctuation. The exp ( E/@) law satisfies these con- 
ditions, and is laid down by Gibbs on grounds of mathe- 
matical convenience, and because of its ready application to 
practical problems, qualifications which are illustrated 
freely in his book. Of course, if we choose to regard the 
individual systems as huge molecules and imagine them to 
be all immersed in an enormous body of gaseous fluid (a 
" temperature -bath "), we could appeal to the Maxwell- 
Boltzmann result and prove that the systems would, except 
on rare occasions, be distributed in phase, as Gibbs laid 
down for his canonical ensemble ; and this explains why in 



ENSEMBLES OF SYSTEMS. II 265 

recent literature, references to Gibbs' canonical ensemble 
as a " temperature -bath " occasionally occur. Indeed, in 
Chapters xx. to xxin., this point of view has been 
actually adopted at certain parts of the argument. 

Of course, the expression exp ( E/0) is multiplied by a 
constant factor in the specification of a canonical system, 
and Gibbs writes his rule in the form 

... . (24.5.2) 



where ^ is a constant for every system, just as @ is. In 
fact, ^ is a function of & and the parameters of a system, 
the parameters, as already stated, being any quantities 
necessary to express unchanging metrical properties of a 
system or its dynamical relations to external bodies which, 
one must remember, are the same for all the systems at one 
moment. is determined in terms of @ and the para- 
meters a v . . . a e by the equation 



all phases. 

or 

>j, / (* E 

e~^= . . . e~i^dq l . . . dp ft . (24.5.3). 

all phases. 

Gibbs devotes several pages to the discussion of average 
values in a canonical ensemble. Thus if u is a function of 
co-ordinates, momenta and parameters, its average value 
over the whole ensemble is given by the expression 

q, . . . dp, . (24.5.4). 



all phases. 

As an illustration let us work out the average value of 

3E 

p r . where p r is any momentum. It is 
dp r 

3E 

P ^ r 

all phases. 



266 STATISTICAL MECHANICS FOR STUDENTS 

Now by (24 .5.3) thia is equal to 

f* f 3E 

J ' ' ' J Pr to 

J ~" W* 



f f 
' ' ' 

J og J 



. (24.5.5). 



But by integration by parts 



f * 

Jj 



Hence (24 .5.5) becomes simply 0. Thus the average 
value of p r d~E/dp r for any momentum is the same for all. 

Now E = Ep -f E ff where E^ is the kinetic energy and 
E ? is the potential. E ? is independent of the momenta, 
and Ep is a homogeneous quadratic function of the momenta. 
So 



I 

2 r =l Pr ty f 
I 



This is Gibbs' version of the equipartition of energy 
theorem. The kinetic energy of any system can be divided 
into /J parts, and the average value of any part over the 
whole ensemble is the same, viz., -| 0. Note that this 
theorem is not found for a time -aver age of a part for a 
single system, but for an average at one instant over all the 
systems of the ensemble. 

Finally we shall show one of Gibbs' thermodynamic 
analogies. In equation (24 .5.3) let us vary the modulus 
and parameters from the values 0, a v . . . a e to + 80, 
a l + 8a v . . . a e + Sa e . We obtain 



* 

e~ 



ENSEMBLES OF SYSTEMS. II 267 

or 

exp 



(24 . 5 . 6). 

Now the expressions 9E/9a s represent the forces 
exerted by the system on the external bodies ; for since E p 
does not depend on the parameters which define the position 
of the external bodies, and since the energy entering the 
system through the action of external bodies is 

z ^ &,. 

,=i da s 

ZSa. 

s~\ca 8 

it follows that the expressions 8E/3a 5 represent the forces 
acting on the system due to external bodies. Writing A, 
for 9E/3a g , i.e., the force exerted by the system on its 
environment through the parameter a 8 , we can write 
(24 . 5 . 6) in the form 

8^^~ E 8fe> - 2 A, 80, . . (24.5.7) 



where we recall (24 .5.4) and indicate an average value of 
quantity over the ensemble by drawing a bar over it. 

Writing <>(#, p, a) for the function (^P E)/@ and 
4> (a) for (9 E)/@ we have 

e __ 

8V = o 80 E A, 8a s 

=i 
and since 



it follows that - * - 

8E + 27 A, 8a 8 . . (24 . 5 . 8). 



268 STATISTICAL MECHANICS FOR STUDENTS 

The analogy with the entropy equation in thermodynamics 
is apparent. <t> (a) is the analogue of entropy, E(a) of the 
internal energy, 2A 8 8a s of the work done by a thermo- 
dynamic system on its environment, & of the temperature, 
M* (a) of the free energy. 

NOTE ON THE INTEGRATION IN SECTION 24 . 4 

If we integrate dg x d 2 throughout a region of plane space 
defined by x 2 + f 2 2 < r 2 the result is Trr* 2 . We obtain 
from this result the integral 'of d^ d 2 d 3 throughout a 
region of space defined by ^ 2 -f- 2 2 + 3 2 <; r 2 by means of 
the " zone " method. That is, we put it equal to 

2 ' 77 (r 2 - x 2 ) dx 



If we now write V n (r) for the result of integrating d^ 
dg 2 ...... dg n throughout a range of values defined by 

i 2 + ^ 2 2 + - . . +&<** 

we can obtain V w (r) from V ;i _ x (r) in an analogous manner. 
Thus 



f- 
= 2r f V n _i ( 

Jo 



r cos cos 



where we write r sin < for x. 

Putting V n (r) = A n r w , where A n is some numerical 
multiplier depending on n, we have 



The definite integral is known to have the value 

1 . 3 . 5 . . . (n 1) 77 ., . 

------------------- v ------- / __ ii n is even, 

2 .4.6 ... n 2 



ENSEMBLES OF SYSTEMS. II 269 

and 2 . 4 . 6 . . . (n - 1) .. . , , 
i if n is odd. 

1.3.5... n 

So 



A 1 . 3 . 5 . . . (n 1) TT A .. . 

A n = 2 . A M T if n is even 

2.4.6 ... n 2 n ~ 1 



and 

A , 2. 4.6.. .tt 1. ., . , , 

A n = 2 . A-., if n is odd. 

1.3.5... n l 

In either case A n = A n _ 2 , 
n 

and thus since A 2 = TT and A 3 = - 7 , it is not hard to show 

o 

that 



V n (r) = r n if w, is even 



n-l 



and - 2 - - - r n if n is odd. 

1 . 3 . 5 ... ft 

Taking w as even, the region within a " shell " bounded by 



and f ! 2 + n 2 = ( r 

is given by SV n (r), or 

n 

U 7T 2 

2 



(i)' 

which is equal to 



In the text r 2 is 2m (E aw 6e), and Sr 2 or 2r Sr is 
2m SE, while n/2 is 3a/2 or c. 



APPENDIX ON RECENT DEVELOPMENTS 

QUITE recently, no later than 1924 in fact, an unexpected 
turn was given to the statistical theory of physical systems 
by the publication of a paper by Bose on Planck's radiation 
formula. The novelty lay in Bose's modification of those 
steps in statistical theory which are concerned with d priori 
probability. Yet so far is it true to say there is " nothing 
new under the sun " that we can find the germ of Bose's 
idea in Planck's earliest papers on black body radiation, 
and it will prove serviceable to refresh one's memory on the 
discussion in section 14 . 2. In the traditional methods of 
statistical theory each particle (molecule, atom, electron) 
in a system is assumed to have a recognisable individuality, 
and in this way one complexion is distinguished from another 
even if they are embraced in the same state of numerical 
distribution of the particles between the phase-cells. In 
section 14 . 2, however, we considered complexions as settled 
by the distribution of elements of energy among the particles, 
and this led to the substitution of an expression such as 

(n + c - 1) ! 
n ! (c - 1) ! 
for the value of a probability W instead of the usual 



n \ n 2 ! . . . n c ! 

It is expressions of the former type which enter into 
Bose's analysis and link it up with Planck's earlier treatment. 

1. Bose's Statistics of Light Quanta in a Temperature-En- 
closure. In his work Bose regards radiation as composed of 
light-quanta, or particles with energy hv where v is the 
frequency of the quantum. In accordance with present 
views as to matter and energy, a particle is regarded as 
having mass Jivjc 1 * and momentum hv/c where c is the 

* I.e., mass at velocity c ; their " rest-mass " is zero, according to 
Relativity theory. 

270 



APPENDIX ON RECENT DEVELOPMENTS 271 

velocity of light. Each light-quantum is represented in a 
six-dimensional phase-diagram representing position (x, 
y, z) and momentum (f , 77, ). Of course 

i 2 + i? 2 + e = (Av/c) 2 . 

If we integrate dx dy dx d df] d throughout the region of 
the phase -diagram in which the energy is not greater than 
liv where v is a particular frequency we obtain 

477 



V - 



3 



where v is the volume of the enclosure containing the radia- 
tion. In this region of phase-extension no points represent- 
ing quanta of higher frequency than v can occur. 
In the extension 



will occur these points which represent quanta with fre- 
quencies between v and v + Sv. If we now introduce the 
assumption that the elementary phase-cells have the 
magnitude h 3 , then the number of the cells in the shell 
bounded by the frequencies v and v + v is 

4 77 V 



The problem is to determine the distribution of the num- 
bers of light-quanta among the various frequencies in an 
enclosure of volume v containing full radiation, i.e., radiation 
in temperature equilibrium with the walls of the enclosure. 
This condition determines the amount of energy in the 
radiation, but it does not determine the total number of 
light-quanta present. We cannot assign an unchanging 
individuality to any light-quantum in the enclosure, since 
the walls are emitting and absorbing radiation ; even the 
number in the radiation is not necessarily unchangeable ; 
for if a quantum of frequency v is absorbed by the wall it 
can be replaced by emission of several quanta of frequencies, 
v', v", . . . without any breach of the energy condition, 
provided v is equal to the sum of v ', v" . . . . Conceive the 



272 STATISTICAL MECHANICS FOR STUDENTS 

whole range of frequencies to be divided into elementary 
ranges between the frequencies v v i> 2 , . . . v r , etc. . . . 
where we can write 8v r for v r v r _- L . Suppose the total 
number of light quanta in the radiation at any moment to 
be n, and that of these n^ lie in the range o to v l9 n% in the 
range v l to v 2 , . . . n r in the range v r ^ l to v r , etc. We 
have to find the number of complexions which corresponds 
to this numerical distribution among the elementary ranges. 
It is here that Bose resorts to the earlier method of Planck. 
Take, for instance, the n r quanta in the range v r _ l to v r * 
Their representative points lie in a shell of the phase- 
diagram which is constituted of a r elementary phase-cells 
where 

/T ,. 

Vr r ' ' ' ' ( ^ 



As we have seen in section 14.2 there are 
(n r + a r - 1) ! 



(1.2) 



n r \(a r -\)\ 

different ways of partitioning the n r light-quanta among 
the cells when we disregard individuality of the quanta. 
The formula, in fact, goes back to (2 . 5) as it gives the 
number of actual terms in the expansion of 



where for the moment we are writing 6 for a r and p for n r . 
The coefficient 

p\ 



Pi\p 2 i. . .p b \ 

of a term such as x^ 1 x 2 p * . . . x b p t> is disregarded. It 
entered into the earlier work because we assumed that each 
particle had a recognisable individuality. Now we regard 
the term x-f 1 x 2 Vt . . . x^ just as representing one way of 
partitioning of p light-quanta among the 6 cells, and there 
are just (p + b 1) \/[p ! (6 1) ! ] such terms, i.e., ways 
of partitioning. 

Expressions similar to (1 . 2) give the number of ways of 
partitioning the n l light-quanta among the a l cells of th$ 



APPENDIX ON RECENT DEVELOPMENTS 273 

range o to v v the /i 2 light-quanta among the a 2 cells of range 
v l to i> 2 > an d so on Any combination of special ways of 
partitioning in the separate shells gives one way of partition- 
ing the total n light-quanta over the whole phase-diagram. 
Hence the total number of ways of distributing n light- 
quanta is on this view of the matter equal to the product 
of the terms such as (1.2); i.e., it is 





n r \ (a, - 1) ! 
where S n r = n, 

and each a r is defined by (I . 1). 

There is, of course, the energy condition to be satisfied, 
viz., 

Zn r Uv r = ^ ..... (1.4) 

where E is a constant ; but there is no constant n condition. 
The problem is now to find the values of n v n 2 , . . . n r , . . . 
which make (I . 3) a maximum subject to the one condition 
(1 . 4). The procedure is the familiar one. Calling (I . 3) 
W (n, a), we take its logarithm, use Stirling's theorem, 
and apply the variational method. 
So we have to solve 
8{Z (n r + a r ) log (n r + a r ) n r log n r a r log a r } = 0* 

subject to 

v r %n r = 0. 

Thus using an undetermined multiplier a we easily find 
that for each value of the integer r 

log (n r + a r ) log n r av r = 
and so n r + a r 



n r 



. . 
= exp (av r ) 



r exp (av r ) 1 (I 5) 

Equations (I.I) and (1 . 4) will then determine a as a 
function of E and v ; for (I.I) shows that 

4 77 V V? 8v r 

n r = - - . 

c 3 exp (av r ) 1* 

* We can drop out minus unity as negligible. 



274 STATISTICAL MECHANICS FOR STUDENTS 
It follows that 

v c 3 exp (av r ) 1 

or going to the limit we find for the energy density 




exp (av) 1 



In this reasoning we have omitted, so far, any reference 
to one point which will be familiar to those acquainted with 
optical theory, and which has made its appearance in these 
pages in Chapter xxn. It is the matter of polarisation 
of light. The effect of allowing for this is to double the 
above result and make it 




exp (av) I 

To find a in terms of the temperature 6 of the enclosure 
we introduce the usual expression for the entropy, viz., 
k log W m , where 

log W m =Z [K, + a r) log (n r + a r ) - n r log n r - a, log a r ] 

the n r having the values given by (1.5). If we now raise 
the temperature the energy of the radiation increases to 
E + SE, and there is a corresponding increase of entropy 
from S to S + SS. The limit of 8S/SE is 0" 1 , since the 
volume v of the enclosure is maintained constant, and no 
external work is done by the pressure of the radiation. 
Writing e r for liv r and //, for a/h we have 

S = TcZ { (n r + a r ) log (n> + a r ) - n r log n, - a r log 



r +a r )(logn r + av r ) n r log n r a r 
= kza r (log n r - loga r ) + (n r + a r ) ^ 
= k 2 [ a r log [exp (p r ) 1] + a, pe r +n r 
r (/ie r - log [exp (^ f ) - 1]) j + A/n E 

- Za r log [1 - exp (- ^ r )] } . 



APPENDIX ON RECENT DEVELOPMENTS 275 

So it follows that 
= ^ 

, , , _, du. , ( d log fl exp ( ,)] ) 

= kp + lcE-- -kS\ a r 8 L j v ^' n 

cufj f dfj j 

du, j ( a r . . du, ] 

^ L ^ ] ' : tXV ( fJL r ) v - f 



exp (p, r ) - 



and thus JJL has its usual value, viz., (kQ)~ l . So we arrive 
ultimately at Planck's expression for the energy-density 
in the temperature-enclosure, E/v, viz.. 



exp (hv/k0) 1- 

If. Einstein's Theory of an Ideal Gas. The method of 
Bose is therefore justified by its success. Einstein immedi- 
ately applied the same method to the discussion of a mona- 
tomic gas. The region of the phase-diagram in which will 
fall all the points representing any molecule whose energy 
does not exceed the value e is given by integrating dx dy 
dz d dt] d( throughout a range of values determined by the 
volume v of the vessel and the condition 



Its magnitude is 



3 

The region in which lie the points corresponding to mole- 
cules whose energies lie between e and e + Se is thus 



and the number of elementary phase-cells in this region is 
thus 

2771; (2m)* 

<* *< 



276 STATISTICAL MECHANICS FOR STUDENTS 

As before, we divide the phase-diagram into shells by the 
energy-hypersurfaces corresponding to e v 2 , . . . e r , . . . 
and write e r - r _ 1 Se r . The number of cells in the shell 
bounded by r _ l and e r is a r where 

a^Vse, .... (II. 1). 



The partitioning is carried out as above, and the number of 
complexions corresponding to an energy content of the gas 
equal to E is W (n, a), i.e., expression (1 . 3) where 

S n r r = E. 

But, of course, we have now the condition to satisfy that 
Z n r n constant, 

for although we disregard individuality of molecules in 
counting complexions, we must remember that in this case 
no particles are created or destroyed. When we take the 
logarithm of W(n, a), vary it, put the variation equal to 
zero and take account of 

Z r Sn r = 

and E 8n r 

we obtain 

log (n r + a, r ) log n r = A + /**,, 
involving two undetermined constants as usual. Thus 

. . (II. 2) 



exp (A + ii r ) 1 

where, of course, a r is given by (II . 1). 

The entropy S is k log W m where W m is the value of 
W(n, a) with this expression for n r inserted. It is easily 
shown as above that it is given by 

S = k [n\ + /xE - Za r log [1 - exp (- A - p, r )] } 

(II . 3). 

On working out dS/dEt as above, we find, as before, that 
the constant ^ is connected with the temperature 6 by the 
usual result. Putting in the values for the a r it follows, 



APPENDIX ON RECENT DEVELOPMENTS 277 

after a step or two, that the number of molecules whose 
energies lie in the range e to -f Se is given by 

^Sc ... (II. 4) 

where the constant A is determined by the condition 

2 TTV (2m)' 
n v ' 




The result (II . 4) replaces the Maxwell Boltzmann law, 
to which it clearly reduces in case where unity is negligible 
in comparison with exp(X + jue). Further it easily follows 
that 




_ 5) 

8 - 

%/0 

and 

Q ^, E 2 TTV (2 m)*k C , . ,_ x weN , 

S = RA + ~ - - ^ > I * log (1 - c-*-*') de 

^ (II. 6). 

III.' The Fermi-Dirac Statistics. This change in the 
manner of counting complexions was suggested, as we have 
said, just on the eve of the third period in the history of 
quantum ideas. About the same time a suggestion was 
made by Pauli in connection with the Bohr theory of atom 
structure, which is called Pauli's exclusion principle ; it 
concerns the description of the quantum paths of the 
electrons in an atom in terms of quantum numbers, and 
postulates that in a given atom or molecule no two electrons 
can have the same set of quantum numbers, i.e., that at any 
instant two electrons cannot be on the same quantum 
orbit. In statistical theory we have seen already the 
analogy of quantum paths of representative points and 
phase-cells, and in 1926 Fermi suggested that in dealing 
with the statistics of systems of molecules we should reckon 
up the complexions on an analogous exclusion principle, 
viz., that complexions in which two or more representative 
points occupied one energy cell should be ruled out as 
impossible. Independently, Dirac made the same sugges- 



278 STATISTICAL MECHANICS FOR STUDENTS 

tion, and also showed how it could be regarded as a reason- 
able conclusion from the new quantum mechanics, which 
was just about a year old at the time, and which had also 
produced a theoretical justification of the original Pauli 
exclusion principle concerning electrons in atoms. The 
effect of this on the formulae of statistical theory is not 
hard to discover. Like Bose and Einstein, Fermi and 
Dirac abandon any idea of individualising molecules, but 
concentrate, as it were, on the individuality of the elementary 
phase-cells in an " energy-shell " of the phase-diagram. 
For Bose and Einstein the occurrence of any number of 
representative points in a cell was as likely as the occurrence 
of any other ; for Fermi and Dirac this is not so ; a cell 
can contain one point or none ; each alternative is equally 
likely ; but it is impossible that it should contain two or 
more. The earlier results of Section II. are simply repeated, 
viz., 

2 TTV (2m)* i ,> /TTT .. x 

a ' = ^r~ L ^ 8c, . . . Ill . i 

A 3 

where E n r e r = E 

and Z n r = n. 

But instead of taking W(n, a) as given by (1 . 3), it is 
determined by 

W(n,o) = n a ' 1 . . (III. 2) 

n r I (a r - n r ) \ 

for the number of ways of arranging n things in a boxes on 
the understanding that no box is to hold more than one 
thing is just the number of combinations of a things taken 
n at a time. The algebraic procedure now follows similar 
lines, and produces these results 

n r = ?L . . (III. 3) 

cp (A +/,) + 1 

S - k log W m 

(- A -/,)]} 

(III . 4) 

and, dS/dE being put equal to 0" 1 , we find the usual result 



APPENDIX ON RECENT DEVELOPMENTS 279 

for jii. The Fermi-Dirac law of distribution replacing the 
Maxwell-Boltzmann law now turns out to be 



with A determined by 

2<rrv(2m)U * , 
n = A_ L ...... dc. 




Also 

O -_/,, /O /wi\>s / ,5 

de . (Ill . 5) 

Jo 

and 
S = RA + ? 




(III. 6). 

The change from minus unity to plus unity is, of course, 
a negligible affair under conditions where the Maxwell- 
Boltzmann law is a justifiable approximation, but, as we 
shall see in a moment, it is quite otherwise under more 
unusual conditions . But even when the Maxwell-Boltzmann 
law is a good approximation the new statistical theory pro- 
duces the value of the entropy constant in a very direct 
and easy manner. Thus if we neglect the ^ 1 we obtain 

x r i 
n ye M e* e ** ae, 



where y is written for 2 TTV (2w)^/& 3 , and so writing # 2 for 
fjie and using the familiar results for i # f e"""* 1 dx we find 

Jo 



n - ye~ T 

so that e x = ~ (2 TTW, k9)% . . . (Ill . 7). 

nh 3 

For hydrogen under normal conditions this gives 6 X 10 4 
for A, and so justifies the approximation used, but if v or 6 
are very small, i.e., at large densities or low temperatures, A 
is too small to permit the neglect of the unity term. Pro- 



280 STATISTICAL MECHANICS FOR STUDENTS 
ceeding, however, under the suitable conditions we find that 
E'=ye~ x r^e-^dc 

Jo 



1 E 

and so 



n 2 

the familiar energy-partition law. 

Hence S = RA + ? + kye~ K f c* eT* dc, 

u Jo 

remembering that log (1 x) x if x is sufficiently 
small compared to unity. Thus 



S = RA + + Jen 

u 



III . 7 = R flog v log n 3 log h +- log 
L Q r,-i 2 



by 



T? T 5 i / i , 5 i i (2 7rm/t) 3 A;~| 

- R [-log 6 - log p + - + log-L pJ J 

(III. 8) 
where p is written for ^A;0 as usual. 

To decide between the two results one must consider 
conditions under which the unity term is not negligible, 
and then experimental evidence leans towards the Fermi- 
Dirac result, and so supports the conclusions of the new 
quantum mechanics. The trouble with the Einstein result 
is that while it makes the entropy of a gas tend to zero as 
the temperature decreases to zero (thus extending the 
Planck-Nernst heat law to all substances, condensed or 
not), it makes the specific heat first rise to a maximum 
and then decrease to zero asymptotically. Now this 
maximum property has not been observed for any gas, and 



APPENDIX ON RECENT DEVELOPMENTS 281 

so the Einstein-Bose statistical theory lacks support in that 
particular. On the other hand, with the Fermi-Dirac 
statistics, it can be shown that at low temperatures 

E = E + a 2 

where ^ __ 3 /6n\*nh* 

~~ 40 (in?) ~m 

j o/^" V \* m 9979 

and a = 2 ( ) - ?r 2 n 2 k 2 

\9n/ rc& 2 



so that the specific heat 2 aO approaches zero as a limit 
without any intervening maximum. It is clear also that 
the Fermi-Dirac statistics necessitates a nul-point energy ; 
this is evident from the fundamental postulate, for of all 
the molecules whose representative points fall within a 
group of phase-cells which have one volume-element in 
common only one can be in that phase-cell of the group 
which corresponds to zero velocity. All the others will 
fall into cells with as small energy as possible, but up 
to a certain limit each cell will have one point in it. In 
the Einstein-Bose statistics the constant A is essentially 
positive, and has zero as its limiting value ; this is obvious 
from result (II . 4) ; otherwise n r would be negative for 
small enough values of e. At the limit A all the repre- 
sentative points would crowd into the cell of lowest energy, 
since l/(e^ e 1) is infinite for the zero value of e. This 
is an extreme case of what is called " degeneracy." So as A 
decreases, either with decreasing temperature or increasing 
density, the system passes from the classical Maxwell- 
Boltzmann distribution to a state of extreme degeneracy ; 
and for small values of A particles with a low velocity are 
present in greater numbers than is the case of a classical 
distribution with the same density and temperature. In 
the Fermi-Dirac statistics, on the other hand, A may have 
positive or negative values.* The distribution differs 
markedly from the classical when v (2 7rmkd) ? ~/(nh*) is much 
less than unity ; A is then large and negative, thus making 

* N.B. The values of A are not necessarily the same for the three cases, 
even when density and temperature are the same. 



282 STATISTICAL MECHANICS FOR STUDENTS 



small compared with unity, except for very large 
values of e. This means that for a large range of e, n r = a r 
(see III . 3), or every cell has one representative point in it. 
The system is degenerate. In cases of degeneracy the 
Fermi-Dirac statistics implies that molecules with low 
velocities occur in greater number than in a classical dis- 
tribution at the same temperature and density. 

The reader should for further information consult a 
paper by Lennard-Jones in the Proc. Phys. Soc. (London), 
Vol. XL., Part 5, (August, 1928), where a full bibliography 
of papers on these matters will also be found. 

IX, The Statistical Method of Darwin and Fowler. Parti- 
tion Functions. An innovation of a different character has 
been introduced lately by Darwin and Fowler. The 
original papers are to be found in the Phil. Mag., Vol. 
XLIV., Nos. 261, 263 (September, November, 1922). 
They begin by pointing out the weak spot in the practical 
working out of the statistical calculations on the usual 
lines, viz., the use of Stirling's Theorem. This point is not 
always brought home in current accounts. Let us, for 
example, consider a gram-molecule of a gas under standard 
conditions with a volume about 20 litres. The energy is 
1*5 R0, about 3 x 1C 10 ergs. If this were all in one molecule 
its momentum would be (1-5 Rra0)*, about 10~ 7 grammem, 
for hydrogen. Thus the whole momentum extension 
would be of the order 1C"" 20 in the phase-diagram, and when 
associated with the volume 20 litres, the phase-extension 
of the phase-diagram actually required for a gram-molecule 
of hydrogen at ordinary temperatures and pressures would 
be of the order 10"~ 16 (erg-sec.) 3 . Suppose this is now divided 
into elementary cells, h* each having, therefore, an extension 
of the order 10~ 79 . There are 10 83 such cells in the whole 
extension. But there are only about 10 24 representative 
points to partition among them ! But the use of Stirling's 
theorem quietly assumes that in practice the points avail- 
able are much more numerous than the cells. 

Of course, on purely classical lines we can retort that we 
are not compelled, in order to arrive at the classical result, 
to postulate cells as small as Planck's element. Still the 



APPENDIX ON RECENT DEVELOPMENTS 283 

use of that element in quantum discussions of the " classical- 
quantum " type requires subsequent justification of the 
results. 

Darwin and Fowler show that this rather illegitimate use 
of Stirling's theorem can be avoided entirely by considering 
the average state of a system, rather than its " most prob- 
able/' as the analogue of the state of thermodynamic 
equilibrium. In practice this leads to no new result, but 
the point of view and method are quite different. We can 
illustrate this by a simple example. Suppose we have 
present a set of Planck vibrators which each contain a 
multiple of a unit of energy . Let there be n vibrators 
and an amount of energy E available where E/e is an 
integer. Let us consider a complexion in which there are 
n vibrators with no energy, % with , n% with 2e, etc. 
There are as usual 

n 1 
n 1 n^ \n 2 \ . . . 

complexions in a corresponding statistical state, and we 
have to satisfy the conditions 

n + n l + n 2 + n 3 + = n Ty 

1 + 2tt a + 3n 8 + . . . =E/ l ' '' 

Let C be the total number of complexions compatible 
with these conditions. C will be given by 

nl 

C = : : i 

n ! % ! n 2 ! . . . 

where the summation is over all possible values of the n 
which satisfy (IV .1). If we consider the series 

(1+2 + z 26 + z 36 + . . .) n . . (IV . 2) 
the typical term is 



where the n r satisfy (IV .1). Thus if we pick out the 
coefficient of 2 E in the expansion (IV . 2) we shall have 
the value of C. So C is the coefficient of Z E in (1 z )"~ n , 



284 STATISTICAL MECHANICS FOR STUDENTS 

i.e., of y in (1 y)~ n where c = E/e. But by the binomial 
theorem 



1. Z. 1. Z. o. 

(^_+c_-^! 
^ (w - 1) ! c ! y ^ 
and so p __ (n + c 1) ! 

~~ (n- l)!c! 

a result which we have already met several times. 

The average value of any quantity denoted by u can 
then be determined by the equation 



Cu =2 _ _ uz (n l + 2n t + zn, + ...)< m (IV . 3). 
n I n 1 ! n 2 I . . . 

Thus the average value of the energy is given by 



, 
n \ n \ 



<fe v ' 

d I 

z 



dz (1 z') n 



2 ~ log (1 - z )l (IV . 4). 
dz J 

The expression (IV . 4) is obtained so as to render possible 
a method of approximate calculation which discards Stirling's 
theorem and uses the method of " contour integration." * 
In this z is regarded as a complex variable, and by integrat- 
ing round a circle in the Argand diagram with the origin as 
centre, and a radius less than unity, we know that 



* See Jefferies' Operational Methods in Physics (Camb. Math. Tracts). 



APPENDIX ON RECENT DEVELOPMENTS 285 
and 




These expressions are exact, but in the theory of contour- 
integrals there is a method known as the " method of 
steepest descents " which leads to simple approximations 
that are adequate and rigorous under certain conditions 
satisfied in most problems. The process is this : 

The integrands being infinite at z = 0, and also at z = 1, 
there must be a minimum value of the integrand at some 
point on the positive axis where z = ft (ft is of course, a 
real number less than unity). Now take the circular 
contour round which the integration is effected to be a circle 
of radius ft. It can be shown that as we travel along this 
circle the values of the integrands have strong maxima 
at z ft and drop to relatively very small values at all 
points on the contour outside the very short arc containing 
the point z = ft. Lest there be any confusion on the part 
of the reader about these statements, let him note that the 
value of an integrand at z = ft is a minimum for all the 
values at points on the real axis between z = and z 1. 
But it is maximum for the values at points along the circle. 
Indeed, if n and E are large enough this maximum is so 
strong that practically the whole of the contour integral is 
contributed by a short arc of the circle in the neighbourhood 
of z = ft. Under these circumstances we can remove the 
term n z(d/dz) log (1 z e ) outside the integral sign provided 
z is given the value ft. 

This value ft is thus obtained by solving the equation 

i.__j = 

7 Tt 1 4- I / i *\v) ' 



or practically 



d l = 0. 



_ _ _ 

dz z E (l 
This equation is easily seen to be 

E (1 z') UZ* = 0, 



286 STATISTICAL MECHANICS FOR STUDENTS 
so that & is given by 



We then have, owing to these approximations 

c = A 



dif 



where A is some constant. 
So that 






Now, of course, this result is entirely trivial in this case. 
As we assumed that the energy of the system is constant, 
obviously the average energy of the system over all the 
complexions is just the constant energy for each complexion, 
as (IV . 6) and (IV . 7) show. The example, however, is 
the easiest illustration of the process of approximation 
employed. Let us choose one which does not lead to a 
trivial result, viz., a system which consists of a vibrators 
with a fundamental quantum energy e, and 6 vibrators with 
a fundamental quantum 77. For the purposes of the calcula- 
tion we must assume that e and 77 are commensurable, but 
it does not matter how large are the integers which are 
required to express 6/77 in its lowest terms. We may, of 
course, choose the unit of energy so that e and 77 are them- 
selves integers. The number of complexions embraced in 
a state in which a r vibrators of the first type have an 
energy re and b g vibrators of the second type have an 
energy 577 is 



APPENDIX ON RECENT DEVELOPMENTS 287 

where we must satisfy the conditions 

Ea r = a 
Sb 8 =b 
S (ra r + sb 9 r)) = E . . . . (IV . 9). 

The total number of complexions possible with the energy 
E is as before 

C=27 _ ^ __ _ ^ _ (IV. 10) 
ajaja,!. . . 6 !6 X !6 2 !. . . V ' 

where the summation is over all values of the a r and 6, 
compatible with (IV . 9). Now the whole energy E can be 
considered as made up of two parts, E a and E 6 , one the 
energy in all the vibrators of the first type, the other in 
those of the second type. The average value of E a over all 
the complexions is given by 

- K + 2a 2 + 30 3 + . . .) . a ! b I 

- 



2 E n 



As before, we can easily show that C is the coefficient of 

1 



(1 - Z e ) (1 - Z") 6 

and C E is the coefficient of S E in 

1 d 1 

z 



(1 z' 1 )" dz (I z') a 

d_ 
or 1 dz 



(1 z"f (1 z') a 

This leads to the following contour integrals for C and 



CE fl 



= 1 t_dz 1 

~'2m}~it & + 1 '(l*Y(l - 



dz fflz ^ 



2m Z E + 1 (1 - z') a (1 - 



288 STATISTICAL MECHANICS FOR STUDENTS 

There is a strong maximum for the integrands in a small 
arc at z = ft of the contour circle of radius ft where ft is 
determined by 

d 1 ; = o 



(IV. 12). 

v ' 



dz 2 E + 1 (1 z e ) a (1 
so that & is found from the equation 



This leads to the approximation 

A 



C 



ft e ) a (1 

d 



^-a- a E+1 (l-ftT(l-; 

leading to 



9--'- 1 
similarly, 



~V 1/1 Q.cN 

(IV . 13) 



This determines the average partition of the energy 
between the two types of oscillators. The method can 
obviously be extended to any number of different types. 
We have to identify the quantity & with the usual properties 
of the system. It is fairly evident that it is connected with 
the temperature, and if we take the temperature on the 
absolute thermodynamic scale to be the same as that 
defined by the equipartition law for systems following the 
classical dynamical principles, we can find what 0- is. Thus 
suppose that the second type of vibrators have a low 
frequency and a small value of 77. In the limit they follow 
the classical laws, and ^"^ is nearly unity. If it is a little 
greater than unity 

log [i + (#-' - i)] == a-* _ i 



APPENDIX ON RECENT DEVELOPMENTS 289 

or $- 1 = TJ log S-, 

so that t\ _._ 1 

" 



-'-! " log 
If ft" 11 is a little less than unity 

log [1 - (1 - -')] = -(1 - *-'), 

and the same result follows. Thus in the limit the particles 
of a system obeying classical laws have each an average 
energy, log (!/&) This is, of course, kinetic and potential, 
and is &0, so that 

=e- fj - 

where ,u is (&0)" 1 , and we arrive at the usual Planck law of 
partition among the oscillators of the first type, viz., 

E - a 

a 



A generalisation to more general types can be easily 
made. Suppose that particles of type 1, a in number, can 
be in states for which the energies are e 1? e 2 , e 3 , etc. ; particles 
of type 2, b in number, can be in states for which the energies 
are rj ly r) 2 , ^ 3 , . . . and so on. Further, we may suppose 
that several quantum states belong to one energy state, so 
that the states of type 1 particles have weights p l9 p 2 , 
p 3 , . . . respectively, of type 2 particles, q l9 q 2 , g 3 , . . . 
respectively, and so on. The weighted number of com- 
plexions which correspond to an assigned specification is 

a\ ai fli 6! bi , 

where 



The functions 

j(z) =p 1 z\+p 2 z e * +Pz#* + - 
+ (z) = qi z* + ?a z" + q 9 z*> + . 
take the place of (1 ?.*)~ 1 9 (1 2 1 )"" 1 , etc., in the previous 



290 STATISTICAL MECHANICS FOR STUDENTS 

example. They are called the " partition functions " of the 
corresponding types of particles. Proceeding as before, we 
find that C is the coefficient of Z B in 

( Z )] [V (2)] . . . 

and C E. is the coefficient of Z E in 



^ft <*)]} [*<*)]* 



i.e., in 

/ j \ 

ft(* 



These are converted into contour integrals, and we have 

ft ()]'[()]*... 



We determine the radius & of the special circle of integration 
by finding the condition for the minimum of the integrand 
in C, i.e., by solving 



dz 



i.e., & practically satisfies 

E = a$ -- log <(&)+&& A log f (&) + ... (IV. 14). 



On putting this value of & into the contour integrals 
above, we again find that approximately 



and (IV . 15) 

<a ft log </> (#) | ft ($)]" 
C E fl = A 



APPENDIX ON RECENT DEVELOPMENTS 291 



so that ^; n d , i /ft \ 

a = a g * ( J 



(IV . 16) 



etc. 

In applying this method to a system containing free 
molecules where quantum states do not strictly enter, we 
can consider the various states defined as usual by elementary 
cells of the phase-diagram 8xSySzSSr)8 * for which the 
energy is ( 2 + r) 2 + C 2 )/2w. The partition function is 
then 



where 8a is written shortly for the sextuple differential 
&g...S, and the suffixes correspond to the various cells ; 
e, is ( r 2 + rj r 2 + r 2 )/2w, and the weight of each state is 
defined to be 8a/h 3 which we have already seen to be the 
justifiable method for combining non-quantum parts with 
quantum parts of a system, on the understanding that 
each quantum state is given the weight unity. In the 
limit x (z) is a sextuple integral, and becomes 



where v is the volume of the vessel. 

Thus suppose there are present a vibrators with partition 
function </> (z) and n free molecules. The parameter 0- is, as 
before, determined by 

E - ab ~ log < () + n -| log x (), 

Ci \T U> t7 

and the energy in the n gas molecules is, on the average, 
equal to 



* N.B. Do not confuse the z co-ordinate of a molecule here with the 
z used as the symbol for the complex variable. 



292 STATISTICAL MECHANICS FOR STUDENTS 

Now 2 <:=:e< logz 



_ log z _ log (1/z) 



where a = 

2m 2m 

400 

SO X () = -^ 



_ (2 77-ra) 3 ; 
~ * [log (l/z)p- 

Hence the energy of the gaseous part of the system is 
equal to 



d ! (2 Trmf* v 

log 



_ 'A 8 [log (I/ft)]* 

ft d i 3 lo 
dftJ2 

371 1 



2 log 

But by the usual elementary result of kinetic theory the 
pressure of the gas is two-thirds of its energy per unit 
volume, and so 

_n I 
P ~v log (I/ft)' 

and since by the usual definition of temperature 

n -, * 

p=-kO, 
v 

we thus arrive once more at the relation between ft and 
temperature 

log (I/ft) -1 =VL 

or ft = e-. 

We can, in a similar manner, work out other mean values. 
Thus the mean over all complexions of the number of 



APPENDIX ON RECENT DEVELOPMENTS 293 

molecules or vibrators of the first type in the r th quantum 
state is obtained thus 

C a, = Za r , "' Pi 1 P Z "' y-n^ - i 6 ' if* . 

' r aj ! a z l . . . b\ b z \ . . . 

= ap r E , ( . a ~ 1 . )! - . Pfpj* 2V'- 1 
*" a>i\ a z \ . . . (a, 1) ! 



where the summation in the second line satisfies the energy 
condition 

a l l + a 2 + . . . + (a r l)<E r + . . . 4 2b t rj t = E e r 
It follows that 



-q^ A 



f),JS - e,- I- 1 

From the value of C in (IV . 15) we find that 



Thus a r is proportional to & er or, e""^ r , and we have the 
usual Planck partition law, degenerating to the Maxwell law 
for gas molecules following the classical laws. 

Another important mean value is the mean of the 
squared energy, E a 2 . Reviewing the argument by which we 
obtained E a , the reader will see that we shall obtain E a 2 if 

we employ the operator (z ) instead of z . Thus 
_ v dz 1 dz 

C E a 2 is the coefficient of 2 E in the expansion of 



and therefore, is equal to 

jfrci]'---, 

~ 



i ( 

2 mJ 



294 STATISTICAL MECHANICS FOR STUDENTS 



which is approximately equal to 



[* (*)]* 



Hence, using the value of C in (IV . 15) we have 



o* 



ftw 

I 

- . . . by (IV . 16) 



The fluctuation of the energy is measured by the mean 
value of the square of the difference E a E a . Now 

(E,- EJ = E a 2- 2 E. Ea + (E,) 2 
and the mean value of this is equal to 



APPENDIX ON RECENT DEVELOPMENTS 295 



dp, 



dd 



de ' 

a result which Einstein made use of in treating fluctuations 
of radiation in a temperature enclosure in the early days of 
quantum theory. The legitimate use of the approxima- 
tion in the case of the operator (z d/dz) 2 requires careful 
analysis, but it appears that it is quite rigorous in a tempera- 
ture bath. 

In the Proc. Eoy. Soc., A 113, p. 432 (1926), will be 
found a paper by Fowler reviewing recent statistical theory, 
and showing how the method of partition functions can be 
applied to the statistics of Einstein and Bose, and that of 
Fermi and Dirac. 



APPENDIX ON COLLISION-FORMULA AND 
CHEMICAL KINETICS 

IN the treatment of systems in statistical equilibrium it is 
postulated that energy can be transferred from molecule to 
molecule, but no assumptions concerning the mechanism by 
which energy is so transferred are required. When, how- 
ever, we deal with systems which are not in equilibrium, it is 
only natural to expect that we shall have to take more 
careful account of the nature of the forces which act between 
molecules ; for while the ultimate state of equilibrium 
attained is independent of the special laws of the intermole- 
cular forces, the manner in which that state is approached 
and the rate of transformation is clearly dependent on them. 

The problems raised by such considerations hardly come 
within the scope of a book which is an " Introduction," and 
which has already grown to the limits set by the capacities 
and needs of its probable readers. However, it may be of 
some value if a further appendix, which will not seriously 
encroach on the student's time and patience, is added, and 
one or two matters which are fundamental in the treatment 
of non-equilibrium states are dealt with. They concern the 
frequency of molecular encounters and the bearing which 
this has on chemical kinetics. 

I. Collisions between Molecules in a Gas. In the Kinetic 
Theory of Gases a very simple type of intermolecular action 
is assumed for many purposes. An encounter (that is the 
interval during which two molecules are within the sphere 
of one another's action) is considered to be so brief in 
relation to the time of free path that it is pictured as a 
" collision " which takes place instantaneously when the 
centres of the two molecules are separated by a definite 
distance which we denote by a. In short, the molecules 
are visualised as " hard spheres," a being equal to the 

296 



APPENDIX ON COLLISION-FORMULA, ETC. 297 

diameter if they are like molecules and to the sum of their 
radii if they are unlike. 

For the reader's guidance through this section, which 
involves some rather tedious steps, it may be as well at the 
outset to give him a preliminary summary to the various 
formulae obtained in it. 

(I. 1) is a general result for the number of collisions per 
unit time between two types of molecules in a gas mixture 
whose velocities are confined to narrow ranges of velocity. 

(I. 2) is the specialised form of (I. 1) when the mixture is 
in statistical equilibrium. 

(I. 3) is the total number of collisions per unit time in the 
mixture when in equilibrium. 

(I. 4) and (I. 5) are formulae for the total number of 
collisions in a gas consisting of one type of molecule. 

(I. 6) is a value for the mean free path of a molecule in a 
gas. 

(I. 7) and (I. 8) refer to the collisions which involve 
a relative velocity between colliding molecules which lies 
within narrow limits or is greater than an assigned value. 

(I. 9) and (I. 10) are similar formulae which involve the 
relative velocity of approach, i.e., the component of relative 
velocity along the line of centres of the colliding molecules. 

The last four formulae are of considerable value in the 
problem of chemical reaction in gases. 

In treating the problem of collision-frequency we consider 
an enclosure in which there are present N^_ molecules of one 
type per unit volume and N% of a second type also per unit 
volume. We introduce a velocity diagram partitioned as 
usual into cells ; at a given instant let the number of 
representative points of molecules the first type in the rth 
cell be v lr Sa> r , where 8aj r is written for 8u r 8v r 8w r (u r , v r) w r 
are the components of the velocity corresponding to the 
central point of the cell). The number of the second type 
is v 2r 8aj r . We are not considering at the moment an 
equilibrium distribution, but there is some law of distribu- 
tion, and so v lr will be some function of u r , v r , w r . Represent 
it by fi (u r , v r , w r , t) ; the time-variable must, of course, be 
involved if the distribution is non-equilibrium. Similarly, 



298 STATISTICAL MECHANICS FOR STUDENTS 

v 2r is equal to/ 2 (u r , v r , w r , t), / x and / 2 not being of necessity 
the same functional form. If the distribution should be ofte 
of statistical equilibrium the function/! (u y v> w, t) would be 
the Maxwell function 



- i pm l (u* + v* + t 

and / 2 (u, v, w, t) would be 

N 2 (^\* exp r / , m2 ( U 2 + V 2 + 

(See expression (4.1.1) and equation (4.1.8), remembering 
that = mu, etc.) Our first problem is to obtain the 
probable number of collisions per unit time between the 
molecules of the first type and those of the second. 

First of all let us analyse the relative situation of two 
molecules, with their representative points in the ath and 
in the 6th velocity-cells respectively, which will lead to a 
collision within an interval 82. Let the reader be careful to 
guard against confusing situations of representative points 
in velocity-cells with situations of the points by which we 
sometimes idealise the molecules themselves in the enclosure. 
The two cells might be very far apart indeed in the velocity- 
diagram without necessarily implying the impossibility of a 
collision within a very brief time of many of the molecules 
represented. It will be realised after little thought that one 
important factor in settling the possibility of a collision 
within an interval St between two assigned molecules is the 
relative velocity u b u a , v b v a , w b w a i another is the 
angle between the direction of this relative velocity and the 
direction of the line of centres of the two molecules from the 
centre of the b molecule to the centre of the a molecule. If 
this angle is denoted by 0, and if or represents the sum of the 
radii of the two molecules, there will be a collision within the 
interval 8t if the centres of the molecules are now separated 
by a distance which lies between a and a + f cos 08t, where r 
is the magnitude of the relative velocity and is equal to 



necessarily less than a right angle.) Thus the centre of any 
molecule of the group b which would collide within time 8 



APPENDIX ON COLLISION-FORMULA, ETC. 299 

with a specified molecule of group a, the " line of centres 
angle " being between and 0+86, would have to lie in a 
ring-shaped space whose circumference is 2 ira sin 0, and 
whose section is an elementary rectangle with sides a80 and 
r cos 08t. The volume of this ring is 2 TrcrV sin cos 80 8t. 
Let us assume that 32 is chosen to have such a value that 
this ring contains on the average one molecule of the group 6, 
so that it is therefore equal to (v 2b 8co b ) ~ l in volume ; then 
8t must be equal to 

1 

v 2b 8aj b 2 770- V sin cos 80 

and in this time there is on the average one collision between 
a specified molecule of the group a and any molecule of the 
group 6 having the defined line-of-centres relation. So in 
one second there are 

v 2b 8oj b 2 7ra 2 r sin cos 80 

collisions between a specified molecule of the group a and 
any molecule of the group 6 having the defined line-of-centres 
relation. Integrating with respect to between the limits 
= o and = 7T/2, we obtain the number of collisions per 
second between a specified molecule of the group a and any 
molecule of the group b. The result is 

V 2b 8a) b 77CT 2 r. 

As there are v la 8a) a of the first type molecules in the ath 
phase-cell, the number of collisions per unit volume per unit 
time between molecules of the first type in the velocity- 
condition represented by the ath velocity cell and molecules 
of the second type in the velocity-condition represented by 
the 6th phase-cell is 

77(72 v la V 2b r & w a 8o>& 

that is, 

2 /i (u a , v a , w a )f 2 (u b , v b , w b ) r ^ (I j. 

du a dv a dw a du b dv b dw b * ' ' 

The total number of collisions per unit volume per second 
between all the molecules of type 1 and all those of type 2 
will be given by the integral of the expression (I, 1) between 
the limits plus and minus infinity for the six velocity com- 



300 STATISTICAL MECHANICS FOR STUDENTS 

ponents. Of course, such an integration could in general 
be carried out only by rather laborious methods of approxi- 
mation. When the gas mixture is in equilibrium and we 
can use the Maxwell law of distribution, the integration can 
be effected by means of the table of intergrals on p. 17, 
though not quite so directly or simply as might appear at the 
first glance. The trouble arises owing to the appearance of 
the factor r, i.e., { (u b u a ) 2 + . . + . . |Mn (I. 1), but the 
difficulty can be surmounted by a transformation of variables 
and, in order to explain this point, we shall have to digress 
for a moment and introduce the reader to two necessary 
lemmas. 

The first is very simple. Let a, j8, y represent the com- 
ponents of the velocity of the centre of mass of two particles 
m I and ra 2 , and let , T?, stand for the components of the 
relative velocity r, so that 



a = 



etc., 

and = u b u a 

etc. 

It is then easy to show that the combined kinetic energy of 
the two particles which is 

i m i ( u a + v <r + w <?) + i 2 ( V + V + *V) 

is equal to 

J (m, + ro a ) (a' + p* + y 2 ) + | m 12 (? + ^+ 2 ) 

= \ ( m i + m*) c * + i m i2 r * 

where c is the velocity of the centre of mass and m 12 = 
m x m 2 l(m l -f- ^2)- This result enables us to write 

fi(u a ,v a ,w a )f z (u b ,v b ,w b )r 
which appears as part of (I. 1) in the form 



when we are considering the equilibrium distribution. 
The second theorem is concerned with substituting for 



APPENDIX ON COLLISION-FORMULAE, ETC. 301 

the sextuple differential du a .... dw h in (I. 1) an expression 
involving the sextuple differential dad^dyd^dr^d^ Con- 
sider for the moment a plane diagram and a point A on 
it which moves over a certain extension in the plane. Let 
x, y stand for the current co-ordinates of A, and let us 
consider another point B whose co-ordinates (X, Y) are 
connected with those of A by the relations 

X = kx + ly 

Y == KX + Xy 

where k, I, K, A, are any constants. The point B will move 
over a second extension in the plane as A moves over the 
first. It can be proved that the area of the second extension 
bears to that of the first the ratio kX K\. This is most 
readily seen by noting that if A 7 and A" are two positions of 
A, and B' and B" are the corresponding positions of B, then 
the area of the triangle OB'B" is equal to 

i (X' Y" - X" Y') 

= i {(**' + ly 9 ) (KX" + Xy") - (Tex" + ly") (KX' + Xy')\ 
= i (iA - K l) (x'y" - x"y') 
= (kX Kl) x area of the triangle OA'A". 

The result stated above follows when one recalls that any 
enclosed area can be divided into elementary triangles having 
a common vertex within the area. 

If we now consider a two-dimensional diagram in which 
we represent u a and u b by a point A, and a and by a point B, 
it follows that an extension in the diagram embracing a 
continuum of values of u a and u b and the extension em- 
bracing the corresponding values of a and are equal in 
area, since in this case 



As this applies to elementary extensions just as much as to 
finite, we can replace du a du b by d&dt; in an integral. A 
similar result holds for the other components. 

After this digression we can return to the general expres- 



302 STATISTICAL MECHANICS FOR STUDENTS 

sion (I. 1), and, in the case of the state of equilibrium, write 
it as 



- | M { (m, + m 2 ) c* + m l2 r*}] r 

da. d^ dy dg drj d . . . (I. 2) 

The integral of this over all possible values of a, /?, y, |, 77, , 
can be separated into two triple integrals, which, apart from 
the initial factors, are 

I ex P ["" 
and 

III 6 ^ [~~ 
By a familiar transformation the first becomes 

f 

4 77 e#p [ | /x (m l + m a ) c 2 ] 

Jo 

the second 

-00 

4 TT I eo:^> [ - J fji m 12 r 2 j r 3 dr 

'o 

The first of these has the value 



W V 2 V 

77 . ^1 1 

4 \fji (mi + m 2 )/ 



(See the table of integrals on p. 17, No. 2.) 
The second has the value 



2 Vju, m lz / 
(See the table, No. 3.) 
Thus the expression (I. 2), when integrated, becomes 



which simplifies to 

. . . (1.3) 



where we recall that \L = (&#) " a , m 12 = m^ m^Km^^ + ^2)? 
and or is the sum of the radii of the two molecules. This 
expression (I. 3) is the total number of collisions per unit 
volume per unit time between molecules of the first kind 



APPENDIX ON COLLISION-FORMULA, ETC. 303 

and molecules of the second kind in a gas mixture which is 
in a state of equilibrium. 

We can easily obtain from this the number of collisions 
in a simple gas whose concentration is N molecules per unit 
volume. The formula (I. 3) will give it if we put N l equal 
to N 2 , m l equal to m a , and divide by 2 ; for, as it stands, 
(I. 3) would count each collision twice. Thus the collision- 
frequency in unit volume of the gas is 

..... (1.4) 

2 A72/ 473r ^M 

or a 2 N 2 ( - ) 2 

\ m J 






where M is the gram-molecular mass, and R the gram- 
molecular gas constant. 

This can be thrown into another very useful form. At 
the end of Chapter IV., we calculated the mean value of 
squared velocities of the molecules in a gas in the equilibrium 
state. We can just as easily calculate the mean velocity. 
If we now use the symbol c to represent the velocity of a 
molecule the number per unit volume whose velocities lie 
between c and c + Sc is 

4 77 N ( )^ exp ( /xrac 2 ) c 2 Sc. 
Hence the average velocity is 

47rf V 2 \exp ( | /Ltmo 2 ) c 3 Sc 

which, by using the third integral in the table on p. 17, is 
equal to 



304 STATISTICAL MECHANICS FOR STUDENTS 

Hence the expression (I. 4) for the collision-frequency in 
unit volume is equal to 



where c is the average molecular velocity. 

Since each collision terminates two free paths there are 
-y/2 7ra*N 2 c free paths described in unit time by the N mole- 
cules. This gives for the average interval of time between 
collisions the value 

1 



V2 77 

and if we multiply this by c we have an estimate for the 
average length of a free path ; it is 

..... (L8) 



There are other methods of calculating mean free path, 
but they all give values approximately to 

7 



Returning to the expression (I. 2) and only integrating 
with respect to a, /?, y, we can obtain the number of collisions 
between molecules of type 1 and molecules of type 2 in a 
mixture whose relative velocity lies between limits r and 
r -f Sr ; it is 

A ** 

4 TT 



~ 

V 477 2 4 

X 4 



that is 

a 2 ^ ^ 2 (2 TTjLAiO* ea:p [- | />tm 12 r 2 ] r 3 Sr . . (I. 7) 

a formula of considerable value in connection with the 
problem of chemical reaction in gases. 

If (I. 7) is integrated for all values of r from a definite 
value from r to oo , we find the number of collisions in which 
the relative velocity is greater than r . Now 



APPENDIX ON COLLISION-FORMULAE, ETC. 305 
f e-""r*dr 

Jr o 

i r 

= -| e-^ydy 

^ J y 

(where y is written for r 2 ) 

1 ( f* f * \ 

= ~ e"(fy- d(e-"*y)[ 

2a (J y^ Jy o } 

(integrating by parts) 



a 



Hence the number of collisions in which the relative velocity 
is greater than r a works out to be 



A 

C 



which by (I. 3) is equal to 

. (1.8) 



where C stands for the total number of collisions per unit 
time. 

It is also of interest in connection with chemical kinetics 
to obtain the number of collisions in which the relative 
velocity of approach of the colliding molecules, i.e., the 
component of r parallel to the line of centres, is within a 
narrow range of values. To do so we revert to the considera- 
tions leading to (I. 1) and (I. 2) and observe, that had we not 
integrated with respect to 6 at an early stage, we would 
have found that the number of collisions between molecules 
with relative velocities between r and r + 8r and line-of- 
centres angle between 9 and + 89 would be given by 
multiplying (I. 7) by 2 sin 9 cos 9 89, i.e., it would be 
2 o- 2 N l N 2 (2 77 ^ 3 m 12 3 )* exp [ J p. m 12 r 2 ] r 3 sin 9 cos 98r89 



306 STATISTICAL MECHANICS FOR STUDENTS 

In such collisions the component of relative velocity parallel 
to the line of centres is practically r cos ; denote it by s. 
If this has to be greater than s , then r must certainly be 
greater than s , and, for a given value of r, cos 6 must lie 
between unity and sjr. So we must integrate the expression 
just written first with respect to 6 between the limits 6 = o 
and = cos ~ * (sjr), and then with respect to r from s to 
infinity. By methods similar to that used for obtaining 
(I. 8) we find the result to be 

C exp [- fji m 12 s*] .... (I. 9) 

If the component of relative velocity is to lie within a 
narrow range of values s to s + 8s, the number is given as a 
differential of (I. 9) ; it is 

2 fjnn lz C exp [ /*w 12 s 2 ] s Ss . . . (I. 10) 

II. Collision-Frequency and Equilibrium. The H Theorem. 
The formula (I. 1) is quite general, and in deriving (I. 2) 
we assumed the Maxwell form for the function / (u, v, w). 
As a matter of fact, the Maxwell distribution law for a gas 
in equilibrium can be deduced from (I. 1), and, indeed, such 
methods of deduction were the first to be employed before 
the work of Boltzmann, Gibbs and Jeans had revealed their 
inherent weakness, and in doing so developed the methods 
of Statistical Mechanics which are " rigorous " in the true 
meaning of that word ; viz., deduced with a clear perception 
of the assumptions which we are making in the proof. A 
few words about this matter will not be out of place. 

Considering the expression (I. 1) for a gas with one type 
of molecule, viz., 

*f ( U a> V a> W a) f K> v b> w b) r d "a da >b> 

we see that if it were integrated for all values of u b , v b , w b > 
between plus and minus infinity, the result would be the 
rate at which molecules whose representative points are in 
the ath velocity-cell are leaving that velocity-state owing 
to encounters. It is theoretically possible, if we know 
enough about the molecules, to determine also the rate at 
which molecules are entering that velocity-state from other 
states. For that purpose it is essential to know sufficient 



APPENDIX ON COLLISION-FORMULA, ETC. 307 

facts about the law of the forces between the molecules. 
As before, the simplest assumption is that of an instantaneous 
collision, combined also with the further assumption of 
perfect elasticity which implies that no energy of translation 
is converted into internal energy, i.e., that the spheres are 
" hard and structureless." 

There is no space to go into details of the proof (the 
reader will find them in full in Jeans' Dynamical Theory 
of Oases, Chapter II.). They show how one can find by 
dynamical methods the equations which give u c , v c , w c , u d , 
v d , w d and ^ in terms of u a , v a , w a , u b , v b , w b and 0, such that a 
collision between a molecule in the cth velocity-cell and one 
in the dth cell with a line-of-centres angle between <f> and 
(f> + 8(f> will result in one molecule entering the velocity- 
state a and the other the velocity-state b, the line-of-centres 
angle lying between 9 and 6 + 89 (with the molecules, of 
course, separating and not approaching). The number of 
such collisions, is, of course, 

277 a 2 f (u c , v c , w c )f(u d , v d , w d ) r' sin c/> cos <S0 8a) c Saj d 

where r' is the relative velocity of molecules of the class c to 
those of the class d. Concerning this expression there are 
two important remarks. First of all, owing to the equations 
referred to above, / (u c , v c , w c ) f (u d , v d , w d ) can be expressed 
as a function of u a , v u , w a , u b , v b , w b and 0. Secondly, 
by an application of the Liouville Theorem, it can be shown 
that the diffcrentical expression r' sin </> cos </> 8(f> Sco c So> (/ can 
be replaced by r sin 9 cos 9 89 8w a 8a) b . It therefore follows 
that the net rate at which molecules in the velocity-state a 
are gaining in number at the expense of other states is 
obtained by integrating the following expression for all 
values of u b , v b , w b , between plus and minus infinity and for 
values of 9 from o to 77/2 

{ / (u c , v c , w c ) f (u d , v d , w d ) - / (u a , v a , w a ) f (u b , v b , w b ) } 
2 Tier 2 8co a r sin 9 cos 9 dw b d9 .... (II. 1) 

Now if the state is one of equilibrium this integral must 
be zero, and one very obvious method of effecting that is 
to make the expression within the { } brackets equal to 

x 2 



308 STATISTICAL MECHANICS FOR STUDENTS 

zero. Proceeding on that line for the moment, it follows 
that 

lo g/ (^a> V a W a) + ^g/ (u b , V b , W b ) 

= log/ (u c , v e , w c ) + log/ (u d , v d , w d ). 

It then appears that the only way to satisfy this functional 
equation consistent with the dynamical relations which 
connect u a , . . . , w b with u c , . . . , w d , is to write the function 
/ (u, v, w) equal to 

A exp [ p{(u u )* + (v ~ v Y + (w w o y\, 

where A, JJL, u , v , w are constants. This is essentially a 
Maxwell distribution for velocities u u o9 v v 09 w W , 
which simply means that it refers to a gas in statistical 
equilibrium whose centre of gravity is moving with a 
uniform velocity u , v , w , in our frame of reference. 

Now from our earlier investigations we know that we 
were bound to reach this conclusion for equilibrium with any 
mechanism which obeys dynamical laws, let alone the very 
simple one we have postulated. Nevertheless, without this 
general support for the argument, one must admit one 
weakness in it. We have, for equilibrium, to equate the 
integral of (II. 1) to zero, and we adopted one very obvious 
way of doing this, but it is by no means obvious that it is the 
only way. It is on general grounds quite possible that 
functional forms of f (u, v, w) might exist which without 
making the integrand zero for all corresponding values of 
the variables would, nevertheless, make the integral of it 
over the whole range of those variables vanish. In that case 
equilibrium would be preserved without the " detailed 
balancing/' as it is called, which would prevail if we make 
the more simple assumption referred to. A great deal of 
discussion on this matter took place during the last years of 
the nineteenth century. At length Boltzmann propounded 
a famous theorem, called the H theorem, which ostensibly 
settled the matter by showing that for elastic collisions, at 
all events, detailed balancing was the only method of 
preserving the gas in its equilibrium state. Yet it was 
pointed out in the discussion that Boltzmann's Theorem 
apparently was in flat contradiction with the reversibility 



APPENDIX ON COLLISION-FORMULAE, ETC. 309 

inherent in dynamical occurrences. In the clearing up 
of this point and in the strict determination of the sense 
in which Boltzmann's Theorem is true, one may say 
that Statistical Mechanics was born and methods for 
the treatment of molecular phenomena without resort 
to any dynamical details of intermolecular collisions de- 
veloped. 

Boltzmann, in his proof of the H theorem began by con- 
structing the function 

f(u, v, w) log/(^, v, w) dudvdw . . . (II. 2) 

the integral extending over all values of the velocity 
components between plus and minus infinity. This is really 
a function of t ; for the reader will remember that in cases 
of non-equilibrium the time-variable should really be 
included among the variables. Boltzmann denoted it by 
the symbol H. If we differentiate the expression with 
regard to the time, we find that 



The value of df/dt for a particular value of u, v } w, can be 
derived from (II. 1) in the manner indicated above, elastic 
collisions being postulated. The details can once more be 
found in Jeans' book. The upshot is to prove that dH/dt 
can only be zero or negative, and that if it is zero, then 
necessarily 

f K> v a , w a ) f (u b , v b , w b ) = f (u c9 v c , w c ) f (u d) v d , w d ) (II. 2) 

Now H depends solely on the law of distribution of the 
velocities at the moment and so remains unchanged if the 
law of distribution remains unchanged. Thus, if there is 
equilibrium it follows that dH/dt must be zero, and it equally 
follows that (II. 2) must be true, and detailed balancing with 
Maxwell's law is deduced as necessary for equilibrium. 
This is satisfactory so far as it goes, but the further part of 
the theorem; viz., that if dH/dt is not zero, it must be 
negative, which appears to be logically bound up with the 
other part, gives us pause. If it is true, then the most 



310 STATISTICAL MECHANICS FOR STUDENTS 

obvious interpretation is that if the state is not an equili- 
brium one, the function H will continually decrease until it 
reaches a minimum value and retain this value thereafter, 
the system having in the meantime attained equilibrium. 
Now this is inconsistent with the dynamical principles from 
which it has been presumably deduced, for by these principles 
any motion is reversible ; so if a certain state of motion is 
possible for the system, that state is also possible in which 
the position of each molecule at any moment is unchanged 
and its velocity exactly reversed in direction. But after 
such reversal, the system, assumed to consist of perfectly 
elastic spherical molecules, would exactly retrace the " path " 
by which it reached this state in the original motion. If in 
that motion H was decreasing, then obviously in this reversed 
motion H will increase, and so to every state of motion in 
which H decreases there corresponds another in which it 
increases. The solution of this paradox is revealed when 
attention is drawn to the fact that in the previous section 
the various formulsD are values for the probable number of 
collisions per unit time ; the predicted behaviour of the H 
function is not given with absolute certainty ; only its most 
probable but not perfectly certain behaviour is given by the 
result. Presumably, if the gas is not in a state of equili- 
brium, H will in all likelihood decrease ultimately to its 
minimum equilibrium value, but it is not guaranteed that it 
may not in the meantime fluctuate now and then to higher 
values than that possessed at the moment. In fact, the 
reader will " see the light " when he recalls the standpoint 
of Statistical Mechanics as he has learnt it in the text, 
especially in Chapters XXIII. and XXIV, We can make no 
definite prediction about the behaviour of the system in any 
given state ; the conditions are too complex to work out in 
detail. We can say with some assurance what is its most 
likely behaviour, our ideas of probability being based on the 
general behaviour of an ensemble of similar systems. Indeed, 
we have in effect met the H function quite early in the book. 
If, instead of using the integral notation of (II. 2), we revert 
to our original notation in connection with the partition of 
a velocity-diagram into c cells, then for a distribution in 



APPENDIX ON COLLISION-FORMULAE, ETC. 311 

which there are n l representative points in the first cell, 
n a in the second, etc., the H function is essentially 



r-l 

apart from a constant term involving the size of the cell. 
Thus H is just 

n log n W (n v ?i 2 , . . . , n c ) 

and the conclusions drawn previously concerning the 
probable but not certain increase of the W function to a 
maximum value are just as true in this strictly limited sense 
concerning the behaviour of the H function. It was, as we 
have stated, in this clarification of the H theorem that the 
true standpoint of Statistical Mechanics, freed, as far as 
equilibrium conditions are concerned, from all connection 
with special collisional mechanisms, stood revealed. 

III. The Kinetics of Gas Reactions in a Homogeneous 
System. If in a mixture of two gases a reaction takes place 
in which the two dissimilar molecules unite to form a new 
molecule or exchange parts to yield two different molecules, 
it is clear that the process is dependent on collisions in some 
way. If, however, a reaction occurs in a simple gas in 
which the original molecules decompose into two or more 
molecules, it is not so obvious that collisions necessarily play 
any part in the process. 

Taking the first case, there is very clear evidence that the 
rate of the reaction, where this is slow enough to be measured, 
is not simply dependent on the collision-frequency, i.e., 
simply proportional to it. Of course, there does not exist a 
state of equilibrium if a chemical reaction is going on ; 
nevertheless, the formulae of Section I. for the equilibrium 
state will apply approximately if the rate is slow enough, 
and if the rate were proportional to the total number of 
collisions per second between the two types of reacting 
molecules, then the rate would be equal to 

KC.C, 

where K is the so-called " velocity-constant " and C l and <7 a 
are the concentrations of the molecules. Now, undoubtedly, 
quite a number of gas reactions occur in which the rate is 



312 STATISTICAL MECHANICS FOR STUDENTS 

proportional to the product of the concentrations ; and 
there exist simple gas reactions in which the rate is pro- 
portional to the square of the concentrations. Yet a glance 
at (I. 3) or (I. 4) will show that on such an assumption the 
velocity-constant would vary as 6* where 6 is the absolute 
temperature at which the reaction is allowed to take place ; 
but this is violently at variance with the facts ; for in such 
reactions of the second order as have been carefully studied 
the value of K will double or even treble for a rise of tempera- 
ture comparable with 10. 

As every physical chemist knows, it was Arrhenius who 
was the fiist to offer a hypothesis to deal with this dis- 
crepancy. About 1890 he suggested that the reaction did 
not take place between molecules in their normal state, but 
between molecules in an " activated " state, i.e., that the 
change was due to collisions between special groups of 
molecules, and this hypothesis fits very nicely into the 
" quantum scheme of things " now existing. For whatever 
might be the change that is produced in the structure of a 
molecule by " activation," it was postulated that it was 
effected by the acquisition of energy far above the normal 
amount in an average molecule, so that nowadays we simply 
regard an activated molecule as a molecule in a higher 
quantum state. For a full discussion of the idea of activa- 
tion, we must refer the reader to a text of Physical Chemistry. 
(See, for instance, W. C. Lewis's System of Physical 
CJiemistnj, Vol. I., Chapter IX. ; Vol. III., Chapter VII.) 
All that we are concerned with at the moment is to show 
how on statistical grounds it leads very simply to Arrhenius' 
well-known law of the change of velocity constant with 
temperature. It is quite sufficient for our purpose to assume 
that each molecule has, as regards internal phases, a normal 
or lower quantum state indicated by the suffix n, and one 
upper quantum state (activated) indicated by the suffix a. 
The result is just as easily proved for the assumption of many 
higher quantum states, but the mathematical expressions are 
more complicated to handle. 

Thus, for the molecules of type 1, the number in the lower 
quantum state are proportional to w ln exp ( p e lw ) where 



APPENDIX ON COLLISION-FORMULA, ETC. 313 

w ln is the a priori probability of that state and e ln the 
internal energy. The number in the activated state is 
proportional to w la exp ( p e la ). (Observe that we are 
assuming the expressions for the equilibrium state to be 
sufficiently good approximations for a state of reaction ; 
this will require some consideration later.) For the second 
molecule similar expressions hold. Thus the number of 
activated molecules of type 1, existing at the moment 
when there are altogether N l molecules of this type which 
have not yet gone into reaction, is 



W 



la 



The activated molecules of type 2 are 



Arrhenius' assumption is that the rate of reaction is pro- 
portional in the collision-frequency of the activated molecules. 
It is not necessarily assumed that all such collisions lead to a 
reaction ; other circumstances, such as state of orientation, 
for example, might have to be taken into account ; but it 
is assumed that a definite fraction of such collisions are 
effective. Calling this fraction / 12 , we see that the rate of 
reaction taken to be the rate at which molecules pass out of 
the reactant condition, viz., -~ d NJdt ord N 2 /dt, is obtained 
as the product of (III. 1), (III. 2),/ 12 and 2 a 2 (27r//zm 12 )*. 
It is further assumed that / 12 does not depend on the 
temperature, being a purely molecular property. Writing 

*_N 1 _dNt_ 

dt & "*"* 

we can easily obtain the expression for K, and on taking the 
logarithm we find it to be equal to the sum of the following 
seven expressions 



log {2a 2 27r/&m 12 } 



, t - a 

log (w ln exp ( ij, e lH ) + w la exp (- p e la )[ 
- log [w 2n exp (- /* ta ) + w 2n exp (- /* e 2n )} 



314 STATISTICAL MECHANICS FOR STUDENTS 

If we proceed to differentiate log K with respect to 0, we 
find that the first two expressions above contribute nothing, 
and d log K/dO is equal to the sum of five expressions 

1 
2~0 

la 



W 



ln 



_ 

k e*{w ln exp(- p, e ln ) + w la exp (- p lfl )} 

and a similar expression for type 2 molecules. 

A little thought will show that the fourth of these is 
simply lm /k 2 , where e lm is the average energy of all the 
molecules of type 1. In effect, e lm is but little different 
from c ln , if the latter is much less than e la . 

To sum up, we find that 



d e 



where is written for (t la + e 2a ) (e lm + e 2m ), i.e., the 
amount by which the combined internal energies of the 
activated molecules exceeds their combined energies in the 
average state, or practically in the lower quantum un- 
activated state. This clearly corresponds to Arrhenius' 
" energy of activation," or as it is sometimes called the 
" critical increment " of energy. Practically %k 6 is 
insignificant compared to , and we obtain Arrhenius' well- 
known equation 

d log K: 

/7 ft If #2 * 

\JU \J K U 

d^_ J_ 

C/JL 7 /\ " 1C j f\n 



As is easily shown, this is quite consistent with the very 
rapid increase of K with temperature. Indeed (III. 3) is 



APPENDIX ON COLLISION-FORMULA, ETC. 315 

used in connection with the experimental determinations of K 
over wide ranges of temperature to determine the value of . 
For details the reader is once more referred to texts of 
Physical Chemistry. 

Despite the rather clumsy appearance of the mathematical 
expressions involved, and the still more complicated expres- 
sions employed if we had used several upper quantum states 
instead of one, the Arrhenius result is essentially dependent 
on the factors of the type e~ M in the formulae for the numbers 
of activated molecules, and it is apparently very satisfactory 
that the result should follow so directly from this universal 
characteristic of statistical formulae. But there is a very 
important feature of the treatment which has to be dealt 
with, and which has in certain cases presented very serious 
difficulties. 

The reaction removes the activated molecules from the 
original system ; they form molecules of the resultant 
substances in an upper quantum state ; their excess energy 
is lost in subsequent collisions (presumably), and they 
become resultants in a normal state, the algebraic difference 
between the energy of activation and the excess energy of 
resultants subsequently lost being the ordinary heat of 
reaction (positive or negative) per molecular group. Of 
course, we know that among the remaining unactivated 
molecules a redistribution of internal energies would take 
place, leading once more to the usual statistical arrangement, 
provided the chemical reaction did not go on. It is clear, 
therefore, that in order to use equilibrium formulae, even for 
approximate calculations, we must be satisfied that the 
reaction goes so slowly that there is always in existence a 
group of activated molecules not too small in number 
compared with the equilibrium number, and that implies 
that the rate of production of activated molecules is at least 
equal to the rate at which they are removed by the reaction. 
Here, then, we are faced with a problem of mechanism. 
We know that the system when denuded of activated 
molecules will proceed to make them good. But will it 
proceed at a fast enough rate? Mechanisms differ, as we 
pointed out above, not in their ultimate goal, but in the rate 



316 STATISTICAL MECHANICS FOR STUDENTS 

at which they reach it. Let us consider the mechanism of 
simple collisions in this connection. On p. 300 there is 
quoted a well-known formula for the combined kinetic 
energies of translation of two molecules approaching a 
collision. If the collision were such as to destroy the 
relative motion so that the molecules travelled thereafter 
with a common velocity (the collision being thus completely 
inelastic) that velocity would by the principle of conserva- 
tion of momentum be c, and the kinetic energy would be 
| (m l + m 2 ) c 2 . Thus \ m 12 r 2 , the " relative kinetic energy " 
before collision, would be converted to some extent into 
internal energies of the molecules ; we say " to some 
extent," for part of it might appear as an energy of rotation 
of the combined molecules. The point is that even in this 
case of a collision without rebound, ^ ra 12 ?' 2 gives the upper 
limit of original kinetic energy available for subsequent 
internal energy, i.e., for activation ; and in other collisions 
even less would be available. If, therefore, we consider the 
rate at which molecules with a definite lower limit of relative 
kinetic energy come into collision with one another, we have 
some means of estimating if the collision-mechanism is 
suitable for a sufficient rapid supply of activated molecules 
whose energy of activation is equal to the assigned lower 
limit. The necessary formulae have been worked out in 
Section I. 

Putting for the energy of activation, formula (I. 8) 
shows us that the number of collisions with a relative 
kinetic energy greater than is 

-^) .... (III. 4) 



Hinshelwood's book, The Kinetics of Chemical Change 
in Gaseous Systems, provides a number of experimental 
results to test this formula. The decomposition of hydrogen 
iodide is a familiar example of a second order reaction, and 
we assume for the moment that the activation is due to 
collisions of normal molecules of HI and the decomposition 
into H 2 and I 2 due to collisions of activated molecules. The 
heat of activation per gram-molecule is 44,000 calories, and, 
taking the gas constant per gram molecule as 2 calories per 



APPENDIX ON COLLISION-FORMULAE, ETC. 317 

degree, this gives 22,000 as the value for /&. The reaction 
has been studied over the range 550 K to 780 K ; so, 
putting as approximately 600, we see that /* (i.e., /&0) is 
over 30, and in comparison unity is negligible in the second 
factor of (III. 4). The formula for C is given in (I. 4), and 
so the rate of collision for molecules with relative kinetic 
energy above is practically 

.... (HI. 5) 

In this we can put a = 2 x 10~ 8 , m = 210 X 10~ 24 , 
lc = 1-35 X 10' 16 , /& = 22,000, 9 = 600, N = 6 X 10 20 
(i.e., a concentration of 1 gram-molecule per litre). The 
result is approximately 

2 X 10 14 

On the other hand, the observed rate of reaction shows 
that about 

2 X 10 13 

molecules of HI (at a concentration of 1 gram-molecule per 
litre and a temperature, 600 K) react per second. So there 
would appear to be a " factor of safety " of about 10, which 
does not appear to be too much to spare, since only a 
favourably circumstanced fraction of the 2 x 10 14 collisions 
per second can result in activation. But we must not over- 
look one assumption which we are implicitly making at the 
moment. It is that all the spare kinetic energy is going 
into one of the colliding molecules, or that about one-tenth 
of all the collisions worth considering result each in one 
activated molecule. Indeed, the treatment so far is tanta- 
mount to assuming that one-tenth of a certain group of 
suitable collisions result at once in the reaction ; for a little 
thought will show that the intermediate state of activation 
and subsequent collision between two activated molecules, 
each one arising from a different collision, is an unnecessary 
part of the picture as we have been treating it so far. In- 
deed, if we consider such a simple collision theory of reaction 
in which activation is not required (or at most is merged 



318 STATISTICAL MECHANICS FOR STUDENTS 

into the process of collision), we can write for the rate of 
reaction 

N* xa* ( ~V e~* expert) (HI- 8) 



where a is an average value for the fraction of collisions 
which result in reaction. As matters stand now, we would 
have to take about 1/10 for the value of a in the case of 
hydrogen iodide, and, in other cases, the fraction does not 
appear to be any less. However, there are other considera- 
tions bearing on internal energy which tend to show this 
simple collision theory of reaction in a more favourable light. 
Before considering these, however, let us glance for a 
moment at a more genuine activation theory, i.e., one in 
which collisions leading to activation and collisions of acti- 
vated molecules leading to reaction are distinct occurrences. 

Looking back to our earlier expressions, we see that the 
energy of activation is the sum of e la e lm and e 2a e 2w , 
so that it is not necessary to assume that an activating 
collision must give to either of the molecules the whole of 
the energy of activation. In the case of hydrogen iodide, 
for instance, an activating collision is only required to give 
one of the molecules engaging in it half the erergy of activa- 
tion ; in a subsequent collision with another molecule 
activated with the half amount, the total amount is then 
presumably available for the splitting into H 2 and I 2 . This 
means that in using (III. 5) for the calculation of the upper 
limit of the number of collisions which might result in 
activation, we do not write exp ( 22,000/0) for exp ( //,), 
but exp ( 11,000/0), which is practically 18 times as great, 
and this certainly offers an ample margin for favourable 
collisions. 

Thus, while there is a good deal to be said on statistical 
grounds for a theory which would consider that bimolecular 
reactions occur through the collisions of previously activated 
molecules, there must be an element of doubt about a simple 
collision theory in which we would assume that normal 
molecules can react at once if they collide with sufficient 
relative velocity. Indeed, if we calculate, as Tolman does, 
on a more stringent basis, which assumes that the kinetic 



APPENDIX ON COLLISION-FORMULA, ETC. 319 

energy corresponding to the resolved component of the 
relative velocity along the line of centres is the only available 
source for the reaction, then even the factor of safety of 10 
or thereabouts disappears, and we have, in general, just a 
rough equality between rate of reaction and rate of favour- 
able collisions. 

However, on other grounds, the theory of activation does 
not seem to be in such favour nowadays among physical 
chemists, and R. N. Fowler has pointed out that the simple 
collision theory can be rendered much more plausible by 
introducing available internal energy considerations, as well 
as merely available relative kinetic energy. In deriving 
Arrhenius' expression, we assumed for simplicity of writing 
just two quantum states, a normal and an activated ; but it 
is much more probable that in reality there are a number of 
quantum states between the normal and the activated. 
Hence, in considering collisions with a definite relative 
kinetic energy, we can assume that some of these will be 
between molecules whose internal states, while not high 
enough for activation, are not as low as the normal. These 
would contain internal energy available for reaction pur- 
poses, and would obviously allow us to put the necessary 
amount of relative kinetic energy at a lower figure than 
before, and, in effect, bring into the favourable field many 
collisions previously ruled out. To render the necessary 
mathematical analysis which develops this idea as simple as 
possible, we will revert to the picture of internal harmonic 
oscillators as the seat of this internal energy, and carry out 
the calculations along classical lines. The internal energy is 
(for / oscillators) 



If we wish to calculate the number of molecules whose 
internal energy lies between 77 and 77 + 877, we must integrate 



for all values of q l . . . . p f which correspond to 



320 STATISTICAL MECHANICS FOR STUDENTS 

divide the result by the same integral over all values of 
q v . . . , p fy and multiply by N. The method of carrying 
out such integrations as occur in the numerator is associated 
with the name of Dirichlet, and will be found in standard 
texts of mathematical analysis.* The result sought for is 
known to be 



f 00 

1 x f ~ l exp ( JJLX) dx 

Jo 

The denominator is, of course, equal to 

1 f 00 

~ y f ~ 1 exp(y)dy 

PJo 



and the integral is known to be equal to (/ 1) ! 

Hence the number of molecules whose internal energy lies 
between 77 and 77 -f- 877 is 

NfL/ r-i-n (in. 7) 



(/-I)! ' 

Using (I. 7) and (III. 7), we can now calculate the number 
of collisions per unit volume per unit time between molecules 
of type 1 with internal energy between 77 1 and 77 x + 877!, and 
molecules of type 2 with internal energy between 772 and 
772 + 8772, the relative kinetic energy being between and 
| -f- S. It is the product of 



2 a* -L exp (- 

\ >IZ / 

and two expressions of the form (III. 7). The result is 



[ p. (rj l + rj 2 + f )j 87) v 8773, 8^ 

To calculate the number of collisions for which the energy 
available for reaction purposes, viz., ^1 + ^2 + >i g greater 
than an assigned value we must integrate this expression for 
* Whittaker and Watson's Modern Analysis (3rd Edition), p. 238. 



APPENDIX ON COLLISION-FORMULAS, ETC. 321 

all values of rj l9 7? 2 , , satisfying this condition. Dirichlet's 
method is once more used, and yields the result 

ao^y. / 2. V /1+/1+ ,C I+/t+1< .- Mrfr 

(/!+/+ 1) I \ii/ J 



which is equal to 



(A+A+i)! 

If /*, or /kd is sufficiently large, this is equal to 



(A 

If the molecules are of one type the factor 2 must as usual 
be removed, and we obtain 

n2/+ 1 

- exp ( //,) . (III. 8) 



Let us assume then that (III. 8) gives the law of reaction 
, where 



^. 

1)1 

a being some fraction giving on the average the relative 
number of sufficiently energetic collisions which result in a 
reaction ; then, taking the logarithm of K and differentiating, 
we find 

dlogK _ _2/+| 
d 9 k9* 6 



But, experimentally, we find the energy of activation 
from the Arrhenius formula 

dlogK __ ^ 
d0 ~~lc6* 9 

where e indicates the experimental value of the energy of 
activation as distinct from , which represents the lower 
limit of the theoretical available energy. 



322 STATISTICAL MECHANICS FOR STUDENTS 
Hence 



and so (III. 8) becomes, in terms of the experimentally 
determined energy of activation, 



while (III. 5) becomes 



which is just (III. 9) when / is made zero. 
Now (III. 9) bears to (III. 10) a ratio which is practically 



(2/+1)! 

an expression which may easily be greater than unity. 
Thus, for hydrogen iodide, / is at least unity, and the ratio 
becomes if that value is assumed 



<r 2 . 



As e is more than thirty times as great as kO in this case, 
this ratio is about 20, and it follows that to meet the experi- 
mental rate of reaction, the simple collision theory, if 
supplemented by this hypothesis of internal available 
energy, would only require a to have on the average a value 
about 1/200, which leaves a very good margin indeed. 

On statistical grounds, then, the collision theory of bi- 
molecular homogeneous gas reactions can be regarded in a 
favourable light. It is, however, a well-known fact that 
serious difficulty has hitherto attended similar considerations 
when applied to reactions of the first order in which the rate 
of reaction is proportional to the first power of concentration 
and not to its square, the best discussed example being the 
decomposition of nitrogen pentoxide. We can, of course, 
as before, assume that the molecules entering into reaction 
have been previously activated by collision, but it is hard to 
admit that reaction is duQ to collision of activated molecules, 
since the rate of reaction would surely be proportional to 



APPENDIX ON COLLISION-FORMULAE, ETC. 323 

square of concentration and not to first power. In order to 
evade this difficulty, it is generally assumed that once a 
molecule has been activated, there is a definite chance that 
before it can be deactivated by another collision, it will 
spontaneously disintegrate. That is, collisions leading now 
to activation and now to deactivation, maintain an equili- 
brium distribution in the usual statistical manner between 
the normal and various quantum states (including the 
activated state) or nearly so, for combined with this is this 
postulated disintegration mechanism which tends to upset 
this distribution (but not too rapidly) by denuding the 
system of its activated molecules. Since the concentration 
of activated molecules is proportional to exp ( juej, it 
follows that the number of activated molecules at any 
moment is 



w n exp (- ii n ) + w a exp (- pc a ) 

and remains practically unaffected if the rate of reaction is 
slow enough ; as there is a definite chance that such a 
molecule will disintegrate before another collision tends to 
deactivate it, this rate of reaction will be proportional to the 
expression just written and so to the first power of N. 
Moreover, K will be equal to 



w n exp(p, e n ) + w a exp(p, fl ) 

where A is a constant independent of the temperature. Taking 
the logarithm and differentiating with respect to 0, we find 
just as before Arrhenius' equation 

d log K 

dO = kid 2 
where = e a e m , or practically a e n . 

This picture of the occurrence is due to Lindemann 
(Trans. Faraday Soc. y Vol. 17, p. 599 [1921]). The difficulties 
attending it have as usual centred round the necessity of 
finding the rate of activation to be fast enough. It is im- 
possible in this short appendix to go into the matter at any 
length. The case of nitrogen pentoxide proves to be the 
most refractory. Thus if activation had only relative 

Y2 



324 STATISTICAL MECHANICS FOE STUDENTS 

movement to look to for its energy, it can be shown that it 
would take place at a rate about -0001 of that required. 
That is Tolman's view, using his very stringent formula 
involving relative motion resolved along the line of centres 
as the only effective source of energy of activation. (See 
Tolman's Statistical Mechanics, Chapter XXI., section 326.) 
To be sure, if the full relative motion is considered, matters 
do not look so bad, and, as Fowler shows, if we supplement 
the energy of relative motion with internal energy, using a 
reasonable number of internal degrees of freedom, we can 
just bring the case of nitrogen pentoxide within the bounds 
of possibility without anything to spare, however ; more- 
over, there are three or four other unimolecular gaseous 
reactions which are well within the grasp of Lindemann's 
hypothesis, as amplified by Fowler, without making too 
great demands on the probabilities of the situation. (For 
details, consult Fowler's Statistical Mechanics, Chapter XVII. 
section 4.) 

Apart from Fowler's considerations, however, and indeed 
some years before he advanced them, Christiansen and 
Kramers (Zeitschrift fur physik. Chem., Vol. 104, p. 451 
[1923] ) attempted to meet the difficulty of Lindemann's 
hypothesis by pointing out that the resultants of the decom- 
position would contain energy above the normal ; such 
energy must leave them when becoming normal resultant 
molecules, and it seems natural to look for the cause of this 
loss in ordinary collisions which convert this energy into 
heat motion of the molecules. But these authors suggest 
that this energy might be used at all events in part for 
activation purposes. In short, activated molecules of the 
resultants will collide with unactivated molecules of the 
reactant and transfer this energy as internal energy to the 
latter, little or none of it transforming into translatory 
kinetic energy. The difficulties attending this hypothesis 
are discussed by Tolman in the reference cited above. 

Finally, it remains to refer briefly to another theory of 
activation propounded nearly fifteen years ago by W. C. 
Lewis, and independently about the same time by Perrin. 
A full account of it will be found in Lewis's System of Physical 



APPENDIX ON COLLISION-FORMULAE, ETC. 325 

Chemistry, Vol. III., Chapter VII. It looks for the energy 
of activation, not to collisions but to the thermal radiation 
surrounding the reacting molecules, and in temperature 
equilibrium with the walls of the enclosing vessel. It 
assumes that a molecule to become active absorbs a quantum 
of energy whose frequency corresponds to the value given 
by the equation, energy of activation = hv. If one uses the 
expressions for rate of absorption of radiation developed in 
the Planck theory of full radiation, however, this hypothesis 
finds itself in as bad, if not a worse, situation as regards 
nitrogen pentoxide than the simple collision theory of 
activation. The advantage of it is that it gives a simple 
explanation of the unimolecular nature of the reaction 
without resort to hypotheses of spontaneous disintegration. 
As regards bimolecular reactions, the radiation theory is 
quite satisfactory if it is assumed that the radiation has only 
to supply a portion (say one-half) of necessary energy of 
activation to one molecule and the remaining portion to the 
other, collisions between previously activated molecules 
producing the reaction. This radiation theory seems to 
have receded into the background lately, although Tolman 
in his book discusses it in a very favourable way, showing 
how an elaboration of it may be possibly made to fit the 
facts. The author had many private discussions on this 
matter with Professor Lewis some years ago, and has 
attempted to present the simple radiation theory in a form 
which yields a criterion distinguishing molecules engaging 
in unimolecular reactions from those engaging in bimolecular 
and at the same time suggests a possible escape from the 
rate of activation difficulty (Rice, " Note on the Radiation 
Theory of Chemical Reaction/ 7 Trans. Faraday Soc., No. 63, 
Vol. XXI., Part 3). Any reader with a sufficient knowledge 
of electromagnetic theory could follow the argument 
advanced in this note, but the author must point out that 
his reasoning leads to an abnormal reflecting power of 
nitrogen peroxide for radiation of the activating frequency 
which would be just below the visible red, and some un- 
published observations carried out by E. A. Stewardson at 
the author's request do not confirm this. 



326 STATISTICAL MECHANICS FOR STUDENTS 

As matters stand now, it would appear that a collision 
mechanism is the least unsatisfactory hypothesis for activa- 
tion. It is quite possible, in view of the solution of similar 
difficulties of quantum states in other processes by the New 
Mechanics, that the theory of activation, if it survives, will 
look in this direction for a better formulation. 

As regards the application of the formulae of Section I. to 
the treatment of viscosity and diffusion in gases, the reader 
will find a sufficiently elementary account in the English 
translation of Bloch's Kinetic Theory of Gases (published by 
Methuen), and in view of the elegant and simple treatment 
to be found in Chapter III. of that little volume, there 
appears to be no need to continue this appendix further. 



NOTE ON CHAPTER X 

The Smoluchowski formula used in connection with 
fluctuations was derived in Chapter X. from a rather detailed 
consideration of the circumstances, so as to make the argu- 
ment as concrete as possible to the beginner. It can, 
however, be derived in a more abstract fashion by a method 
due to Einstein. 

Consider the normal state of a system in which the 
internal energy is U and the free energy F, so that 

F== U - 6S 

where S is the entropy, or 

U-F 



Indicate a state to which the system can fluctuate by 
primed symbols, the internal energy and temperature still 
being the same however. Let this state be one to which 
the system could be brought from the normal by external 
work of amount S E performed on the system. Then since 

8F = -SS 6+ SE 
and S is zero, it follows that 

F' - F + 3 E 
U-F-SE 



and S = 



SE 



But 8 = k\ogW 

and S' = k log W 



327 



328 STATISTICAL MECHANICS FOR STUDENTS 

by Boltzmann's formula. Hence 

W _ fS' - S 

W = 



or W = W exp 

which is Einstein's general formula. In Chapter X. 

8E = - f" (p - p ) dv 

J v 

and Smoluchowski's result follows. 



SUGGESTIONS FOR FURTHER 
READING 

THOSE readers who wish to pursue the subject further 
mainly for its applications in Physics and Chemistry will 
find an excellent guide in R. C. Tolman's " Statistical 
Mechanics " (American Chemical Society Monograph Series, 
published by the Chemical Catalog Company, New York). 
The recently published " Statistical Mechanics " of R. H. 
Fowler is a very exhaustive treatise and, for those who 
possess sufficient mathematical equipment, a veritable mine 
of information on the many topics, physical, chemical and 
astrophysical, to which the statistical method can be applied. 
It is published by the Cambridge University Press. These 
books give numerous references to the original literature. 

The classical work is, of course, Willard Gibbs' " Elemen- 
tary Principles in Statistical Mechanics " (Yale University 
Press). Based on the treatment of ensembles of similar 
dynamical systems its " modest aim," in the words of its 
author, is that " of deducing the more obvious propositions 
relating to the statistical branch of mechanics." It does 
not in the main concern itself with thermodynamic phe- 
nomena or with the " mysteries of Nature," confining itself 
to logical deduction without reference to hypotheses con- 
cerning the constitution of matter. Nevertheless, no serious 
student of the subject should fail to read it, and its four last 
chapters do at all events discuss the relations of the subject 
to natural phenomena with that insight for which the 
author was so justly famous. 

A very important monograph by P. and T. Ehrenfest 
discusses the origin of the subject and the gradual clarifica- 
tion of its fundamental postulates. Its title is " Begriffliche 
Grundlagen der statistischen Auffassung in der Mechanik." 

329 



330 SUGGESTIONS FOR FURTHER READING 

It is Heft 6 of Band IV. 2 II. of the " Encyclopadie der 
Mathematischen Wissenschaften." 

There are several works on the applications of the statis- 
tical method to gases. The book for those whose time and 
mathematical knowledge are limited is E. Bloch's " Kinetic 
Theory of Gases," a translation of which is published by 
Methuen. The standard works are, of course, Boltzmann's 
" Vorlesungen iiber Gastheorie," and Jeans' " Dynamical 
Theory of Gases." An American work, " Kinetic Theory of 
Gases," by Loeb (Published by Ginn & Co.), combines the 
usual theoretical treatment with useful accounts of recent 
experimental research on gases. 



INDEX 

Action, 79 

Action-integrals, 140 
A priori probability, 6 
Avogadro's hypothesis, 51 

Bohr's postulates, 147 

Bose's statistics of light-quanta, 270 

Characteristic temperatures, 210 

Chemical constant, 215 

Chemical kinetics, 311 

Clapeyroii's equation, 120 

Collision-frequency, 296 

Complexions of a molecular system, 21, 24 

Condensation, 104 

Contour integration applied to statistical calculations, 284 

Darwin and Fowler's statistical method, 282 
Debye's theory of specific heats of solids, 209 
Degeneracy in a conditionally periodic system, 158, 161 
Distribution constant, 47, 49 
, modulus of, 263 

Einstein's fluctuation formula, 327 

statistics of an ideal gas, 275 
theory of specific heat of solids, 207 
Elastic spectrum, 175, 189 
Energy, average, 45 

equipartitioti of, 57, 60, 266 
hypersurfaces, 153, 156 
of vibrating lattice, 181, 195 
Ensemble, canonical, 261 

micro canonical, 263 
of systems, 234, 246 
Entropy, 69 

constant, 80, 213, 221, 258, 279 
kinetic, 75 
of a perfect gas, 77 
Equation of state of a perfect gas, 47 
Equilibrium, chemical, 82 
Error, mean square, 19 
Errors, normal law of, 11, 14 

331 



332 INDEX 

Fermi-Dirac statistics, 277 
Fluctuations, 102 
Fourier's theorem, 175, 176 
Free -energy, 75 
Full radiation, 205, 275 

Gas reactions, 82 

Gibbs' canonical ensemble, 261 

microcanonical ensemble, 263 

phase-space, 254 

thermodynamic analogies, 266 

Hamilton's equations, 246 
H-theorem, 306 

Intermolecular forces, 92 

Lattice, cubical, 123, 189 

energy of, 195 
vibration of, 192 
Lattice, linear, 175 

energy of, 181 

vibration of, 176 
Lattice, superficial, 186 

Lattices, cubical, statistics of a system of, 198 
Lattices, linear, statistics of a system of, 183 
Light-quantum, 141, 270 
Liouville's theorem, 248 

Maxwell-Boltzmann law, 43, 277 
Mean-squared velocity, 56 
Microscopic states, 252 
Mixture of gases, 49 
Modulus of distribution, 263 
Molecular phase-space, 254 
Molecules, finite size of, 92 

Nernst's heat theorem, 213 
Normal state of a system, 68 
Nul -point energy, 141, 281 

Oscillator, 57, 131 

Parameters of a system, 38 

Partition functions of Darwin and Fowler, 282, 290 

Pauli's exclusion principle, Fermi's adaptation of, 277 

Phases, 24 

Phase-cell, 25, 77 

finite magnitude of, 136, 221 
Phase diagram, 26 
Planck's constant, 80, 130, 134 

law of distribution, 134, 137, 141, 156, 273, 293 
for full radiation, 205, 275 



INDEX 333 

Planck-oscillator, 131 
Pressure of a fluid, 53 

, intrinsic or internal, 54, 94, 96 
Pressures, law of partial, 62 
Probability, d priori, 6, 34, 151 

and intermolecular action, 97 

state of maximum, 39 
Propagation of a disturbance in an elastic solid, 190 

Quantisation of paths, 134, 147, 158 
Quantum hypothesis, 126 
states, 136, 138 

Rayleigh-Jeans law for full radiation, 205 
Rotational specific heat, 166 

Saturated vapour, theory of, 118 

Second law of thermodynamics, 68, 114 

Smoluchowski's theory of fluctuations, 106 

Solid state, 123 

Specific heat, 65, 123, 164, 166, 277 

Standing waves in a lattice, 200 

Stationary states, 143 

Steepest descents, method of, 284 

Stern's treatment of the entropy-constant problem, 226 

Stirling's theorem, 24, 28 

Temperature, and distribution constant, 45 

characteristic, 210 
Thermodynamic equilibrium, 68 
Trajectory, 26 

Unstable homogeneous states of a fluid, 96 

Van der Waal's equation, 95, 101 
Vapour -pressure constant, 215 

Wave -function, 189 



PRINTED IN GREAT BRITAIN BY THE WHITEFRIARS PRESS, LTD., 
LONDON AND TONBRIDQE.